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.
Methanosarcina barkeri strain 227 produced ethane during growth on H2/CO2 when ethanol was added to the medium in concentrations of 89–974 mM; ethane production varied from 14 to 38 nmoles per tube (20 ml gas phase, 5.7 ml liquid) with increasing ethanol concentrations. Cells grown to mid-logarithmic phase (A600 0.46, protein = 64 g/ml) on H2/CO2, thoroughly flushed with H2/CO2, then exposed to ethanol, produced maximal ethane levels (at 585 and 974 mM ethanol) of about 215 nmoles per tube, with an ethane/methane ratio of 1×10-3. Mid-logarithmic-phase cultures of Methanosarcina barkeri strain Fusaro also produced ethane (up to 20 nmoles per tube) when exposed to ethanol. Cultures of strain 227 growing on methanol in the absence of H2 produced 6 nmoles per tube of ethane when supplemented with ethanol whereas those lacking ethanol but containing H2 and/or methanol produced 1.6 nmoles per tube. Cultures of Methanococcus deltae strains LH and RC, Methanospirillum hungatei or Methanobacterium thermoautotrophicum produced 5 nmoles ethane per tube when grown in medium containing ethanol. Ethanol concentrations of 177–886 mM were inhibitory to growth of all methanogens examined. Production of ethane by Methanosarcina was inhibited by >62 mM methanol, and both methanogenic inhibitors tested, CCl4 and Br–CH2–CH2–SO inf3 sup- , inhibited ethane and methane production concurrently. The data suggest that ethanol is converted to ethane by Methanosarcina species using the terminal portion of the methanol-to-methane pathway.  相似文献   

2.
Red pine (Pinus resinosa Ait.) and paper birch (Betula papyrifera Marsh.) seedlings exposed to sulfur dioxide produced acetaldehyde and ethanol, and exhibited increased production of ethylene and ethane. Gas chromatographic measurement of head space gas from incubation tubes containing leaves or seedlings was a simple method of simultaneously measuring all four compounds. Increased ethylene production had two phases, a moderate increase from the beginning of the stress period and a large increase just prior to appearance of leaf lesions. Ethane production in SO2-stressed plants did not increase until lesions appeared. Acetaldehyde and ethanol production began within 6 hours at 0.3 microliter per liter SO2 and 24 hours at 0.1 microliter per liter SO2 and continued throughout a 6-day fumigation. Production of acetaldehyde and ethanol continued when plants were removed to clean air for up to 2 days. A higher concentration of SO2 (0.5 microliter per liter) induced acetaldehyde and ethanol production within 2 hours of the start of fumigation of birch and pine seedlings. A number of other stresses, including water deficit, freezing, and ozone exposure induced production of acetaldehyde and ethanol. Production of these compounds was not due to hypoxia, as the O2 partial pressure in the incubation vessels did not decline. Increasing the O2 partial pressure to 300 millimeters Hg did not affect production of these compounds. Production of ethylene, acetaldehyde, and ethanol declined when more than 80% of the leaf area became necrotic, while ethane production was linearly related to the percentage of necrosis. A number of woody and herbaceous plant species produced acetaldehyde and ethanol in response to freezing stress, while others did not. Measurement of these four compounds simultaneously in the gas phase may be a valuable method for monitoring plant stress, particularly air pollution stress.  相似文献   

3.
Methanogenium organophilum, a non-autotrophic methanogen able to use primary and secondary alcohols as hydrogen donors, was grown on ethanol. Per mol of methane formed, 2 mol of ethanol were oxidized to acetate. In crude extract, an NADP+-dependent alcohol dehydrogenase (ADH) with a pH optimum of about 10.0 catalyzed a rapid (5 mol/min·mg protein; 22°C) oxidation of ethanol to acetaldehyde; after prolonged incubation also acetate was detectable. With NAD+ only 2% of the activity was observed. F420 was not reduced. The crude extract also contained F420: NADP+ oxidoreductase (0.45 mol/min·mg protein) that was not active at the pH optimum of ADH. With added acetaldehyde no net reduction of various electron acceptors was measured. However, the acetaldehyde was dismutated to ethanol and acetate by the crude extract. The dismutation was stimulated by NADP+. These findings suggested that not only the dehydrogenation of alcohol but also of aldehyde to acid was coupled to NADP+ reduction. If the reaction was started with acetaldehyde, formed NADPH probably reduced excess aldehyde immediately to ethanol and in this way gave rise to the observed dismutation. Acetate thiokinase activity (0.11 mol/min·mg) but no acetate kinase or phosphotransacetylase activity was observed. It is concluded that during growth on ethanol further oxidation of acetaldehyde does not occur via acetylCoA and acetyl phosphate and hence is not associated with substrate level phosphorylation. The possibility exists that oxidation of both ethanol and acetaldehyde is catalyzed by ADH. Isolation of a Methanobacterium-like strain with ethanol showed that the ability to use primary alcohols also occurs in genera other than Methanogenium.Non-standard abbreviations ADH alcohol dehydrogenase - Ap5ALi3 P1,P5-Di(adenosine-5-)pentaphosphate - DTE dithioerythritol (2,3-dihydroxy-1,4-dithiolbutane) - F420 N-(N-l-lactyl--l-glutamyl)-l-glutamic acid phosphodiester of 7,8-dimethyl-8-hydroxy-5-deazariboflavin-5-phosphate - Mg. Methanogenium - OD578 optical density at 578 nm - PIPES 1,4-piperazine-diethanesulfonic acid - TRICINE N-(2-hydroxy-1,1-bis[hydroxymethyl]methyl)-glycine - Tris 2-amino-2-hydroxy-methylpropane-1,3-diol - U unit (mol substrate/min)  相似文献   

4.
Acetaldehyde strongly binds to the wine preservative SO2 and, on average, causes 50–70 mg l?1 of bound SO2 in red and white wines, respectively. Therefore, a reduction of bound and total SO2 concentrations necessitates knowledge of the factors that affect final acetaldehyde concentrations in wines. This study provides a comprehensive analysis of the acetaldehyde production and degradation kinetics of 26 yeast strains of oenological relevance during alcoholic fermentation in must under controlled anaerobic conditions. Saccharomyces cerevisiae and non-Saccharomyces strains displayed similar metabolic kinetics where acetaldehyde reached an initial peak value at the beginning of fermentations followed by partial reutilization. Quantitatively, the range of values obtained for non-Saccharomyces strains greatly exceeded the variability among the S. cerevisiae strains tested. Non-Saccharomyces strains of the species C. vini, H. anomala, H. uvarum, and M. pulcherrima led to low acetaldehyde residues (<10 mg l?1), while C. stellata, Z. bailii, and, especially, a S. pombe strain led to large residues (24–48 mg l?1). Acetaldehyde residues in S. cerevisiae cultures were intermediate and less dispersed (14–34 mg l?1). Addition of SO2 to Chardonnay must triggered significant increases in acetaldehyde formation and residual acetaldehyde. On average, 0.33 mg of residual acetaldehyde remained per mg of SO2 added to must, corresponding to an increase of 0.47 mg of bound SO2 per mg of SO2 added. This research demonstrates that certain non-Saccharomyces strains display acetaldehyde kinetics that would be suitable to reduce residual acetaldehyde, and hence, bound-SO2 levels in grape wines. The acetaldehyde formation potential may be included as strain selection argument in view of reducing preservative SO2 concentrations.  相似文献   

5.
Summary This paper presents a study of the mechanisms of Cl transport through the brush border membranes of the posterior part of the intestine in the freshwater trout, Oncorhynchus mykiss. The mechanisms for Cl transport in the posterior intestine are distinct from those in the middle intestine; an inwardly directed pH gradient stimulates Cl uptake by bursh border membrane vesicles, indicating a Cl/OH exchange. A pH-regulated Cl conductance is present, which is not activated at normal intracellular pH. Cl uptake is stimulated by an outwardly directed HCO 3 gradient revealing the presence of a Cl/HCO 3 exchange but, conversely, Cl is not exchanged against SO 4 2- . In addition, carbonic anhydrase activities have been detected in both the intracellular and extracellular leaflets of the bursh border membranes which favour the establishment of a bicarbonate gradient. A model of Cl transport mechanisms through the brush-border membranes of the posterior intestine of the freshwater trout is proposed.Abbreviations BBM brush border membrane - CA carbonic anhydrase - EGTA ethylene-bis(oxyethylenenitrilo)tetra-acetic acid - FW fresh water - Hepes N-2-hydroxy-ethyl-piperazine-N'-2-ethanesulphonic acid - Mes 2-(N-morpholino)ethane sulphonic acid - SITS 4-acetamido-4-isothiocyanostilbene-2,2-disulphonic acid - TEA triethanolamine - TMA tetramethylammonium - TRIS tris(hydroxymethyl)aminomethane  相似文献   

6.
Candida glycerinogenes is an aerobe which does not depend on sulphite for production of glycerol. With a sufficient O2 supply, up to 130 g glycerol l–1 was produced with 2.6 g acetic acid l–1 as by-product. However, with an insufficient O2 supply – with higher volumes of medium or at higher corn steep liquid concentrations – the glycerol concentration was lower because the by-products, ethanol, pyruvate and lactic acid, were produced in greater amounts, up to 45 g l–1, 4.3 g l–1, 1.6 g l–1, respectively, whereas, less acetic acid (0.6 g l–1) was produced. In addition, ethanol decreased to 0.4 g l–1 and the glycerol yield improved from 34 to 50% (w/w) by adding 50 g sulphite l–1, nevertheless, acetic acid increased to 7.8 g l–1.  相似文献   

7.
We have studied the ethanolic fermentation of D-xylose with Pachysolen tannophilus in batch cultures. We propose a model to predict variations in D-xylose consumed, and biomass and ethanol produced, in which we include parameters for the specific growth rate, for the consumption of D-xylose and production of ethanol either related or not to growth.The ideal initial pH for ethanol production turned out to be 4.5. At this pH value the net specific growth rate was 0.26 h–1, biomass yield was 0.16 g.g–1, the cell-maintenance coefficient was 0.073 g.g–1.h–1, the parameter for ethanol production non-related to growth was 0.064 g.g–1,h–1 and the maximum ethanol yield was 0.32 g.g–1.List of Symbols A c Carbon atomic weight - a d1/h Specific cell-maintenance rate defined in Eq. (8) - c Mass fraction of carbon in the biomass - E g/l Ethanol concentration - f x Correction factor defined in Eq. (13) - f x Correction factor defined in Eq. (13) - f xi Correction factor defined in Eq. (14) - k d1/h Death constant - M E Ethanol molecular weight - M s Xylose molecular weight - M xi Xylitol molecular weight - m g xylose/g biomass Maintenance coefficient for substrate - m dg xylose/g biomass Maintenance coefficient when k d - q Eg ethanol/g biomass. Specific ethanol production rate - s g/l Residual xylose concentration - s 0 g/l Initial xylose concentration - t h Time - x g/l Biomass concentration - x 0 g/l Initial biomass concentration - Y E/sg ethanol/g xylose Instantaneous ethanol yield - ¯Y E/sg ethanol/g xylose Mean ethanol yield - Y E s/T g ethanol/g xylose Theoretical ethanol yield - Y E s/* g ethanol/g xylose Corrected instantaneous ethanol yield - ¯Y E s/* g ethanol/g xylose Corrected mean ethanol yield - Y x/sg biomass/g xylose Biomass yield - ¯Y xi/sg xylitol/g xylose Mean xylitol yield Greek Letters g ethanol/g biomass Growth-associated product formation parameter - g ethanol/g biomass.h Non-growth-associated product formation parameter - dg ethanol/g biomass.h Non-growth-associated product formation parameter when k d0 - h Variable defined in Eq. (6) or Eq. (7) - 1/h Specific growth rate - m1/h Maximum specific growth rate  相似文献   

8.
A fermentation medium based on millet (Pennisetum typhoides) flour hydrolysate and a four-phase feeding strategy for fed-batch production of baker's yeast,Saccharomyces cerevisiae, are presented. Millet flour was prepared by dry-milling and sieving of whole grain. A 25% (w/v) flour mash was liquefied with a thermostable 1,4--d-glucanohydrolase (EC 3.2.1.1) in the presence of 100 ppm Ca2+, at 80°C, pH 6.1–6.3, for 1 h. The liquefied mash was saccharified with 1,4--d-glucan glucohydrolase (EC 3.2.1.3) at 55°C, pH 5.5, for 2 h. An average of 75% of the flour was hydrolysed and about 82% of the hydrolysate was glucose. The feeding profile, which was based on a model with desired specific growth rate range of 0.18–0.23 h–1, biomass yield coefficient of 0.5 g g–1 and feed substrate concentration of 200 g L–1, was implemented manually using the millet flour hydrolysate in test experiments and glucose feed in control experiments. The fermentation off-gas was analyzed on-line by mass spectrometry for the calculation of carbon dioxide production rate, oxygen up-take rate and the respiratory quotient. Off-line determination of biomass, ethanol and glucose were done, respectively, by dry weight, gas chromatography and spectrophotometry. Cell mass concentrations of 49.9–51.9 g L–1 were achieved in all experiments within 27 h of which the last 15 h were in the fedbatch mode. The average biomass yields for the millet flour and glucose media were 0.48 and 0.49 g g–1, respectively. No significant differences were observed between the dough-leavening activities of the products of the test and the control media and a commercial preparation of instant active dry yeast. Millet flour hydrolysate was established to be a satisfactory low cost replacement for glucose in the production of baking quality yeast.Nomenclature C ox Dissolved oxygen concentration (mg L–1) - CPR Carbon dioxide production rate (mmol h–1) - C s0 Glucose concentration in the feed (g L–1) - C s Substrate concentration in the fermenter (g L–1) - C s.crit Critical substrate concentration (g L–1) - E Ethanol concentration (g L–1) - F s Substrate flow rate (g h–1) - i Sample number (–) - K e Constant in Equation 6 (g L–1) - K o Constant in Equation 7 (mg L–1) - K s Constant in Equation 5 (g L–1) - m Specific maintenance term (h–1) - OUR Oxygen up-take rate (mmol h–1) - q ox Specific oxygen up-take rate (h–1) - q ox.max Maximum specific oxygen up-take rate (h–1) - q p Specific product formation rate (h–1) - q s Specific substrate up-take rate (g g–1 h–1) - q s.max Maximum specific substrate up-take rate (g g–1 h–1) - RQ Respiratory quotient (–) - S Total substrate in the fermenter at timet (g) - S 0 Substrate mass fraction in the feed (g g–1) - t Fermentation time (h) - V Instantaneous volume of the broth in the fermenter (L) - V 0 Starting volume in the fermenter (L) - V si Volume of samplei (L) - x Biomass concentration in the fermenter (g L–1) - X 0 Total amount of initial biomass (g) - X t Total amount of biomass at timet (g) - Y p/s Product yield coefficient on substrate (–) - Y x/e Biomass yield coefficient on ethanol (–) - Y x/s Biomass yield coefficient on substrate (–) Greek letters Moles of carbon per mole of yeast (–) - Moles of hydrogen atom per mole of yeast (–) - Moles of oxygen atom per mole of yeast (–) - Moles of nitrogen atom per mole of yeast (–) - Specific growth rate (h–1) - crit Critical specific growth rate (h–1) - E Specific ethanol up-take rate (h–1) - max.E Maximum specific ethanol up-take rate (h–1)  相似文献   

9.
A possible modulation of permeabilities of membrane vesicles to anions and cations was explored by light scattering techniques, evaluated by measuring the capacity of the vesicles to shrink and swell in response to changes of the osmolarity of the incubation medium. Membrane fractions were obtained by phase partition. Purity was evaluated by detection and quantification of membrane enzyme markers: vanadate-sensitive ATPase for the plasma membrane, nitrate-sensitive ATPase for the tonoplast and azide-sensitive ATPase for mitochondria. Membrane vesicles (250 g protein) were exposed to hypertonic solutions of salts (0.6 osmolar). Kinetics of the changes in apparent absorbance at 546 nm were observed by the addition of potassium, nitrate and chloride salts. The diffusion of ions into vesicles was induced by an osmotic gradient across the membrane and brought about volume changes of vesicles. Upon addition of vesicles to the different solutions the following ion permselectivity sequences were observed: PNO 3 >PCl >PSO 4 2– and PK +PNa +>PNH 4 +.Abbreviations ATP adenosine 5-triphosphate - EDTA ethylene diaminetetraacetic acid - Tris-Mes (Tris[hydroxymethyl]aminomethane, Mes-(2-[N-Morpholino]ethanesulfonic acid) - PEG polyethylene glycol  相似文献   

10.
Granum  Espen  Myklestad  Sverre M. 《Hydrobiologia》2002,477(1-3):155-161
A new method is described for the combined determination of -1,3-glucan and cell wall polysaccharides in diatoms, representing total cellular carbohydrate. The glucan is extracted by 0.05 mol l–1 H2SO4 at 60 °C for 10 min, and the cell wall polysaccharides are subsequently hydrolyzed by 80% H2SO4 at 0–4 °C for 20 h. Each carbohydrate fraction is determined by the phenol-sulphuric acid method. The method has been demonstrated for axenic cultures of the marine diatom Skeletonema costatum and natural marine phytoplankton populations dominated by diatoms. Cellular glucan and cell wall polysaccharides were determined with standard deviations of 1–3% and 2–5%, respectively.  相似文献   

11.
Yeast PAPS reductase: properties and requirements of the purified enzyme   总被引:5,自引:0,他引:5  
The enzymatic mechanism of sulphite formation in Saccharomyces cerevisiae was investigated using a purified 3-phosphoadenylsulphate (PAPS) reductase and thioredoxin. The functionally active protein (MR 80–85 k) is represented by a dimer which reduces 3-phosphoadenylyl sulphate to adenosine-3,5-bisphosphate and free sulphite at a stoichiometry of 1:1. Reduced thioredoxin is required as cosubstrate. Examination of the reaction products showed that free anionic sulphite is formed with no evidence for bound-sulphite(s) as intermediate. V max of the enriched enzyme was 4–7 nmol sulphite · min-1 · mg-1 using the homologous thioredoxin from yeast. The velocity of reaction decreased to 0.4 nmol sulphite · min-1 · mg-1 when heterologous thioredoxin (from Escherichia coli) was used instead. The K m of homologous thioredoxin was 0.6 · 10-6 M, for the heterologous cosubstrate it increased to 1.4 · 10-6 M. The affinity for PAPS remained practically unaffected (K m PAPS: 19 · 10-6 M in the homologous, and 21 · 10-6 M in the heterologous system). From the kinetic data it is concluded that the enzyme followed an ordered mechanism with thioredoxin as first substrate followed by PAPS as the second. Parallel lines in the reciprocal and a common intersect in the Hanes-plots for thioredoxin were seen as indication of a ping-pong (with respect to thioredoxin) uni-bi (with respect to PAPS) mechanism.Abbreviations APS adenylyl sulphate - DTE dithioerythritol - DTT dithiothreitol - HPLC high performance liquid chromatography - IEF isoelectric focusing - LSC liquid scintillation counting - 3,5-PAP adenosine-3,5-bisphosphate - PAPS 3-phosphoadenylyl sulphate - PEP phospho-(enol)pyruvate - SDS-PAGE sodium dodecyl sulphate polyacrylamide gel electrophoresis - Tris 2-amino-2-hydroxymethyl-1,3-propanediol  相似文献   

12.
Summary Unidirectional fluxes of35SO4 across and into rabbit ileal epithelium were measured under short-circuit conditions, mostly at a medium SO4 concentration of 2.4mm. Unidirectional mucosa (m)-to-serosa (s) ands-to-m fluxes (J ms,J sm) were 0.456 and 0.067 moles hr–1 cm–2, respectively.J ms was 2.7 times higher in distal ileum than in mid-jejunum. Ouabain abolished net SO4 transport (J net) by reducingJ ms. Epinephrine, a stimulus of Cl absorption, had no effect on SO4 fluxes. Theophylline, a stimulus of Cl secretion, reducedJ ms without affectingJ sm, causing a 33% reduction inJ net. Other secretory stimuli (8-Br-cAMP, heat-stable enterotoxin, Ca-ionophore A23187) had similar effects. Replacement of all Cl with gluconate markedly reducedJ net through both a decrease inJ ms and an increase inJ sm. The anion-exchange inhibitor, 4-acetoamido-4-isothiocyano-2,2-sulfonic acid stilbene (SITS), when added to the serosal side, reducedJ ms by 94%, nearly abolishingJ net. SITS also decreasedJ sm by 75%. Mucosal SITS (50 m) was ineffective. 4,4-diisothiocyano-2,2-sulfonic acid stilbene (DIDS) had effects similar to SITS but was less potent. Measurements of initial rates of epithelial uptake from the luminal side (J me) revealed the following: (1)J me is a saturable function of medium concentration with aV max of 0.94 moles hr–1 cm–2 and aK 1/2 of 1.3mm; (2) replacing all Na with choline abolishedJ me; (3) replacing all Cl with gluconate increasedJ me by 40%; (4) serosal SITS had no effect onJ me; and (5) stimuli of Cl secretion had no effect onJ me or increased it slightly. Determination of cell SO4 with35SO4 indicated that, at steady-state, the average mucosal concentration is 1.1 mmoles per liter cell water, less than half the medium concentration. Cell SO4 was increased to 3.0mm by adding SITS to the serosal side. Despite net transport rates greater than 1.4 Eq hr–1 cm–2, neither addition of SO4 to the SO4-free medium nor addition of SITS to SO4-containing medium altered short-circuit current. The results suggest that (1) ileal SO4 absorption consists of Na-coupled influx (symport) across the brush border and Cl-coupled efflux (antiport) across the basolateral membrane; (2) the overall process is electrically neutral; (3) the medium-to-cell Cl concentration difference may provide part of the driving force for net SO4 absorption; and (4) since agents affecting Cl fluxes (both absorptive and secretory) have little effect on SO4 fluxes, the mechanisms for their transcellular transports are under separate regulation.  相似文献   

13.
Summary The following equations represent the influence of the ethanol concentration (E) on the specific growth rate of the yeast cells () and on the specific production rate of ethanol () during the reactor filling phase in fed-batch fermentation of sugar-cane blackstrap molasses: = 0 - k · E and v = v 0 · K/(K +E) Nomenclature E ethanol concentration in the aqueous phase of the fermenting medium (g.L–1) - Em value of E when = 0 or = 0 (g.L–1) - F medium feeding rate (L.h–1) - k empirical constant (L.g–1.h–1) - K empirical constant (g.L–1) - Mas mass of TRS added to the, reactor (g) - Mcs mass of consumed TRS (g) - Me mass of ethanol in the aqueous phase of the fermenting medium (g) - Ms mass of TRS in the aqueous phase of the fermenting medium (g) - Mx mass of yeast cells (dry matter) in the fermenting medium (g) - r correlation coefficient - S TRS concentration in the aqueous phase of the fermenting medium (g.L–1) - Sm TRS concentration of the feeding medium (g.L–1) - t time (h) - T temperature (° C) - TRS total reducing sugars calculated as glucose - V volume of the fermenting medium (L) - V0 volume of the inoculum (L) - X yeast cells concentration (dry matter) in the fermenting medium (g.L–1) - filling-up time (h) - specific growth rate of the yeast cells (h–1) - 0 value of when E=0 - specific production rate of ethanol (h–1) - 0 value of when E=0 - density of the yeast cells (g.L–1) - dry matter content of the yeast cells  相似文献   

14.
Summary The ethanol yield was not affected and the ethanol productivity was increased when exponentially decreasing feeding rates were used instead of constant feeding rates in fed batch ethanol fermentations. The influences of the initial sugar feeding rate on the ethanol productivity, on the constant ethanol production rate during the feeding phase and on the initial ethanol production specific rate are represented by Monod-like equations.Nomenclature F reactor feeding rate (L.h–1) - Fo initial reactor feeding rate (L.h–1) - K time constant; see equation (l) (h–1) - ME mass of ethanol in the fermentor (g) - Ms mass of TRS in the fermentor (g) - Mx mass of yeast cells (dry matter) in the fermentor (g) - P ethanol productivity (g.L–1.h–1) - R ethanol constant production rate during the feeding phase (g.h–1) - s standard deviation - So TRS concentration in the feeding mash (g.L–1) - t time (h) - T fermentor filling-up-time (h) - T time necessary to complete the fermentation (h) - TRS total reducing sugars calculated as glucose (g.L–1) - Vo volume of the inoculum (L) - Vf final volume of medium in the fermentor (L) - Xo yeast concentration of the inoculum (dry matter) (g.L–1) - ethanol yield (% of the theoretical value) - initial specific rate of ethanol production (h–1)  相似文献   

15.
Annual, volume-weighted concentrations ofSO4 2– in bulk precipitation have declinedsteadily (–0.44 mol/liter-yr) since 1965 atthe Hubbard Brook Experimental Forest (HBEF), NH inresponse to decreases in regional SO2 emissions(r 2 = 0.74). Similar declines in concentrationshave occurred in wet-only precipitation at HBEF and atnearby sites since 1978. However, decreases inSO4 2– concentrations following passage ofthe U.S. Clean Air Act Amendments in 1990, were notunusual from the perspective of long-term data fromthe HBEF. Statistically significant declines (–5.6mol/ha-yr) in bulk deposition of SO4 2– also have occurred since 1965 in relation to decreases inSO2 emissions (r 2 = 0.58), but annualvariations in deposition also are strongly related toamount of precipitation and other factors.  相似文献   

16.
The effects of UV-B radiation generated in the laboratory and as a component of sunlight on the viability and particular biochemical activities of the bacterium Staphylococcus aureus have been examined. UV-B radiation progressively inhibits protein synthesis (assayed as 3H-alanine incorporation) and kills cells. Cell respiration, and RNA and DNA synthesis (3H-uridine and 3H-thymidine incorporation) were not greatly affected by UV-B irradiation. The OH and 1O2-free radical scavengers protected cells against killing and inhibition of protein synthesis by UV-B, suggesting that such radicals mediate the effects of UV-B on this organism. A similar protective effect using a ferric ion chelator suggests an important role for metallic ions in UV-B lethality.Abbreviations VIS, UV-A, UV-B, UV-C radiation in the bands 400–750 nm, 315–400 nm, 280–315 nm, 200–280 nm respectively - DBCO diazabicyclooctane - OFR oxygen free radical - OH, 1O2, O inf2 sup- hydroxyl free radical, singlet oxygen, superoxide radical respectively  相似文献   

17.
The stability and, consequently, the lifetime of immobilized enzymes (IME) are important factors in practical applications of IME, especially so far as design and operation of the enzyme reactors are concerned. In this paper a model is presented which describes the effect of intraparticle diffusion on time stability behaviour of IME, and which has been verified experimentally by the two-substrate enzymic reaction. As a model reaction the ethanol oxidation catalysed by immobilized yeast alcohol dehydrogenase was chosen. The reaction was performed in the batch-recycle reactor at 303 K and pH-value 8.9, under the conditions of high ethanol concentration and low coenzyme (NAD+) concentration, so that NAD+ was the limiting substrate. The values of the apparent and intrinsic deactivation constant as well as the apparent relative lifetime of the enzyme were calculated.The results show that the diffusional resistance influences the time stability of the IME catalyst and that IME appears to be more stabilized under the larger diffusion resistance.List of Symbols C A, CB, CE mol · m–3 concentration of coenzyme NAD+, ethanol and enzyme, respectively - C p mol · m3 concentration of reaction product NADH - d p mm particle diameter - D eff m2 · s–1 effective volume diffusivity of NAD+ within porous matrix - k d s–1 intrinsic deactivation constant - K A, KA, KB mol · m–3 kinetic constant defined by Eq. (1) - K A x mol · m–3 kinetic constant defined by Eq. (5) - r A mol · m–3 · s–1 intrinsic reaction rate - R m particle radius - R v mol · m–3 · s–1 observed reaction rate per unit volume of immobilized enzyme - t E s enzyme deactivation time - t r s reaction time - V mol · m–3 · s–1 maximum reaction rate in Eq. (1) - V x mol · m–3 · s–1 parameter defined by Eq. (4) - V f m3 total volume of fluid in reactor - w s kg mass of immobilized enzyme bed - factor defined by Eqs. (19) and (20) - kg · m–3 density of immobilized enzyme bed - unstableness factor - effectiveness factor - Thiele modulus - relative half-lifetime of immobilized enzyme Index o values obtained with fresh immobilized enzyme  相似文献   

18.
The pigment composition of two species of green-colored BChl c-containing green sulfur bacteria (Chlorobium limicola and C. chlorovibrioides) and two species of brown-colored BChl e-containing ones (C. phaeobacteroides and C. phaeovibrioides) incubated at different light intensities have been studied. All species responded to the reduction of light intensity from 50 to 1 Einstein(E) m–2 s–1 by an increase in the specific content of light harvesting pigments, bacteriochlorophylls and carotenoids. At critical light intensities (0.5 to 0.1 E m–2 s–1) only brown-colored chlorobia were able to grow, though at low specific rates (0.002 days–1 mg prot–1). High variations in the relative content of farnesyl-bacteriochlorophyll homologues were found, in particular BChl e 1 and BChl e 4, which were tentatively identified as [M, E] and [I, E] BChlF e, respectively. The former was almost completely lost upon reduction of light intensity from 50 to 0.1 E m–2 s–1, whereas the latter increased from 7.2 to 38.4% and from 13.6 to 42.0% in C. phaeobacteroides and C. phaeovibrioides, respectively. This increase in the content of highly alkylated pigment molecules inside the chlorosomes of brown species is interpreted as a physiological mechanism to improve the efficiency of energy transfer towards the reaction center. This study provides some clues for understanding the physiological basis of the adaptation of brown species to extremely low light intensities.Abbreviations BChl bacteriochlorophyll - [M, E] BChlF e 8-methyl, 12-ethyl BChl e, esterified with farnesol (F). Analogously: I - isobutyl - Pr propyl - Car carotenoids - Chlb chlorobactene - HPLC high performance liquid chromatography - Isr isorenieratene - LHP light harvesting pigments - PDA photodiode array detector - RC reaction center - RCH relative content of homologues  相似文献   

19.
The isotope exchange between35S-labeled sulfur compounds of sulfate (SO4 2–), elemental sulfur (S0), polysulfide (Sn 2–), hydrogen sulfide (HS: H2S + HS + S2–), iron sulfide (FeS), and pyrite (FeS2) was studied at pH 7.6 and 20 °C in anoxic, sterile seawater. Isotope exchange was observed between S0, S2 2– HS, and FeS, but not between35S labeled SO4 2– or FeS2 and the other sulfur compounds. Polysulfide mediated the isotope exchange between S0 and bisulfide (HS). The isotope exchange between S0 and Sn 2–) reached 50% of equilibrium within < 2 min while exchange between S2 2– and HS approached equilibrium within 0.5-1 h. In all the experiments HS, revealed a fraction exchange from 0.79 to 1.00. Isotope exchange between S2– and FeS took place only via S2 2– and/or HS. The isotope exchange between iron sulfide and the other sulfur compounds was not complete within 24 h as shown by a fraction exchange of 0.07–0.83. This lack of equilibrium (fraction exchange < 1) was due to the isotope exchange between dissolved compounds and surfaces of sulfur particles. The isotopic exchange reactions limit the usefulness of radiotracers in process studies of the inorganic sulfur species. Exchange reactions will also affect the stable isotope distribution among the sulfur species. The kinetics of the isotopic exchange reactions, however, depend on both pH and temperature.  相似文献   

20.
Thiobacillus tepidarius, isolated from the hot springs at Bath, Avon, UK, grew optimally at 43–45°C and pH 6.0–7.5 on thiosulphate or tetrathionate. In batch culture, thiosulphate was oxidized stoichiometrically to tetrathionate, with a rise in pH. The tetrathionate was then oxidized to sulphate, supporting growth and producing a fall in pH to a minimum of ph 4.8. The organism contained high levels of thiosulphate-oxidizing enzyme, rhodanese and ribulose bisphosphate carboxylase. It was obligately chemolithotrophic and autotrophic. In chemostat culture, T. tepidarius grew autotrophically with the following sole energy-substrates: sulphide, thiosulphate, trithionate, tetrathionate, hexathionate or heptathionate. Thiocyanate, dithionate and sulphite were not used as sole substrates, although sulphite enhanced growth yields in the presence of thiosulphate. Maximum specific growth rate on tetrathionate was 0.44 h-1. True growth yields (Y max) and maintenance coefficients (m) were calculated for sulphide, thiosulphate, trithionate and tetrathionate and observed yields at a single fixed dilution rate compared with those on hexathionate and heptathionate. Mean values for Y max, determined from measurements of absorbance, dry wt, total organic carbon and cell protein, were similar for sulphide, thiosulphate and trithionate (10.9 g dry wt/mol substrate) as expected from their equivalent oxygen consumption for oxidation. Y max for tetrathionate (20.5) and the relative Y o values (as g dry wt/g atom oxygen consumed) for thiosulphate and all four polythionates indicated that substrate level phosphorylation did not contribute significantly to energy conservation. These Y max values were 40–70% higher than any of those previously reported for obligately aerobic thiobacilli. Mean values for m were 6.7 mmol substrate oxidized/g dry wt·h for sulphide, thiosulphate and trithionate, and 2.6 for tetrathionate.Abbreviation PIPES Piperazine-N,N-bis(ethane sulphonic acid)  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