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.
《BBA》1987,893(3):386-397
Three NAD(P)H dehydrogenases were found and purified from a soluble fraction of cells of the purple non-sulfur bacterium Rhodobacter capsulatus, strain B10. Molecular mass of NAD(P)H, NADPH and NADH dehydrogenases are 67 000 (4 · 18 000), 35 000 and 39 000, and the isoelectric points are 4.6, 4.3 and 4.5, respectively. NAD(P)H dehydrogenase is characterized by a higher sensitivity to quinacrine, NADPH dehydrogenase by its sensitivity to p-chloromercuribenzoate and NADH dehydrogenase by its sensitivity to sodium arsenite. In contrast to the other two enzymes, NAD(P)H dehydrogenase is capable of oxidizing NADPH as well as NADH, but the ratio of their oxidation rates depends on the pH. All NAD(P)H dehydrogenases reacted with ferricyanide, 2,6-dichlorophenolindophenol, benzoquinone and naphthoquinone, but did not exhibit transhydrogenase, reductase or oxidase activity. Moreover, NADH dehydrogenase was also capable of reducing FAD and FMN. NAD(P)H and NADH dehydrogenases possessed cytochrome-c reductase activity, which was stimulated by menadione and ubiquinone Q1. The activity of NAD(P)H and NADH dehydrogenases depended on culture-growth conditions. The activity of NAD(P)H dehydrogenase from cells grown under chemoheterotrophic aerobic conditions was the lowest and it increased notably under photoheterotrophic anaerobic conditions upon lactate or malate growth limitation. The activity of NADH dehydrogenase was higher from the cells grown under photoheterotrophic anaerobic conditions upon nitrate growth limitation and under chemoheterotrophic aerobic conditions. NADPH dehydrogenase synthesis dependence on R. capsulatus growth conditions was insignificant.  相似文献   

2.
The nicotinamide adenine dinucleotide dimers (NAD)2 obtained by electrochemical reduction of NAD+ are oxidized by adriamycin in anaerobic photocatalyzed reaction yielding NAD+ and 7-deoxyadriamyci-none. Under the same conditions NADH is not oxidized.  相似文献   

3.
NADH/NAD+ ratios and internal pyruvate concentrations were determined during switches between aerobic and anaerobic steady-state conditions of glucose-limited chemostat cultures of Enterococcus faecalis. During the switch experiments, changes in catabolic fluxes were observed: transition from anaerobic to aerobic conditions resulted in a complete and instantaneous conversion of glucose into acetate and CO2 via the pyruvate dehydrogenase complex, while during a switch from aerobic to anaerobic conditions the culture became homolactic. A similar switch to a homolactic fermentation was observed upon release of the limitation by addition of a glucose pulse to the culture. In sharp contrast to this, a pyruvate pulse resulted in an increase of both pyruvate formate-lyase and pyruvate dehydrogenase complex activity. Furthermore, acetoin was formed during a pyruvate pulse, probably due to a dramatic increase in internal pyruvate concentration. Regulation of the catabolic fluxes over the various pyruvate-catabolizing enzymes is discussed in view of the observed changes in internal pyruvate concentrations and NADH/NAD+ ratios.  相似文献   

4.
Escherichia coli overexpressing a NAD(+)-dependent formate dehydrogenase (FDH) from Candida boidinii was grown in chemostat culture on various carbon sources at 0.05 h(-1) dilution rate, under anaerobic conditions using defined medium and compared to a control without the heterologous FDH pathway. Metabolic fluxes, NADH/NAD(+) ratios and NAD(H/(+)) levels were determined under a range of intracellular NADH availability. The effect of NADH manipulation on the distribution of metabolic fluxes in E. coli was assessed under steady-state conditions. The heterologous FDH pathway converts 1 mol of formate into 1 mol of NADH and carbon dioxide, in contrast with the native FDH where no cofactor involvement is present. Previously, we found that this NADH regeneration system doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed and reached 4.6 mol NADH/mol of substrate when sorbitol was used as a carbon source in a complex medium. In the current study, it was found that higher NADH yields and NADH/NAD(+) ratios were achieved with our in vivo NADH regeneration system compared to a control lacking the new FDH pathway in the three carbon sources (glucose, gluconate and sorbitol) examined suggesting a more reduced intracellular environment. The total NAD(H/(+)) amounts were very similar for all the combinations studied. It was also found that the ethanol to acetate ratio increased with increased NADH availability. This ratio increased from 1.05 for the control strain in glucose to 9.45 for the strain expressing the heterologous NAD(+)-dependent FDH in sorbitol.  相似文献   

5.
Abstract Actinomyces viscosus is a predominant microorganism in dental plaque. It is, just as the oral Streptococcus spp., a saccharolytic and aero-tolerant organism. We have investigated the effects of oxygen on the growth and metabolism of A. viscosus . To this end A. viscosus Ut 2 was grown in a glucose limited chemostat culture on a chemically defined medium ( D = 0.2 h−1) with exposure to variable amounts of oxygen. The Yglucose increased from 62.5 g · mol−1 under anaerobic conditions to 149 g · mol−1 under aerobic conditions, while, concomitantly, the carbon recovery from acidic fermentation products decreased from 75% to 7%. Addition of [14C]glucose to the chemostat showed that the glucose, which was not converted to acidic fermentation products, was instead converted to carbon dioxide or used for the production of biomass. Under aerobic and anaerobic conditions identical cytochrome spectra, containing only two cytochrome b -type absorption bands, were found. It was concluded that electron transport phosphorylation probably occurs both under aerobic and anaerobic conditions. Anaerobically, fumarate served as the electron acceptor, while the high growth yields observed under aerobic conditions are likely to be explained by citric acid cycle activity coupled to electron transport phosphorylation.  相似文献   

6.
A freshwater Pseudomonas sp. was grown in continuous culture under steady-state conditions in L-lactate-, succinate-, glucose- or ammonium-limited media. Under carbon limitation, the NAD(H) (i.e. NAD + NADH) concentration of the organisms increased exponentially from approximately 2 to 7 mumol/g dry wt as the culture dilution rate (D) was decreased from 0.5 to 0.02 h-1. Organisms grown at a given D in any of the carbon-limited media possessed very similar levels of NAD(H). Therefore, under these conditions, cellular NAD(H) was only a function of the culture O and was independent of the nature of the culture carbon source. D had no influence on the NAD(H) content of cells grown under ammonium limitation. In contrast, cellular NADH concentration was not influenced by D in carbon- or ammonium-limited media. In L-lactate-limited medium, bacteria possessed 0.14 mumol NADH/g dry wt; very similar levels were found in organisms grown in the other media. The results are consistent with those of Wimpenny & Firth (1972) that bacteria rigidly maintain a constant NADH level rather than a constant constant NADH: NAD ratio. NADP(H) (i.e. NADP + NADPH) and NADPH levels were also not influenced by changes in the culture carbon source or in D; in L-lactate-limited medium these concentrations were 0.97 and 0.53 mumol/g cell dry wt, respectively. The NADPH:NADP(H) ratio was much higher than the NADH:NAD(H) ratio, averaging 55% in carbon-limited cells.  相似文献   

7.
Abstract: Enterococcus faecalis was grown under anaerobic conditions in chemostat cultures on energy sources with different degress of reduction (i.e. mannitol, glucose, pyruvate) at various culture pH values. Intracellular NADH/NAD ratios were measured and were found to be influenced both by the nature of the energy source and by the culture pH value. Highest ratios were found with mannitol as energy source and with high culture pH values. A role for the redox potential of the NADH/NAD couple as a regulatory effector is suggested by a correlation of the redox potential with the in vivo distribution of the carbon flux between pyruvate formate lyase and the pyruvate dehydrogenase complex.  相似文献   

8.
Role of NAD in regulating the adhE gene of Escherichia coli.   总被引:1,自引:0,他引:1       下载免费PDF全文
The fermentative alcohol dehydrogenase of Escherichia coli is encoded by the adhE gene, which is induced under anaerobic conditions but repressed in air. Previous work suggested that induction of adhE might depend on NADH levels. We therefore directly measured the NAD+ and NADH levels for cultures growing aerobically and anaerobically on a series of carbon sources whose metabolism generates different relative amounts of NADH. Expression of adhE was monitored both by assay of alcohol dehydrogenase activity and by expression of phi(adhE'-lacZ) gene fusions. The expression of the adhE gene correlated with the ratio of NADH to NAD+. The role of NADH in eliciting adhE induction was supported by a variety of treatments known to change the ratio of NADH to NAD+ or alter the total NAD+-plus-NADH pool. Blocking the electron transport chain, either by mutation or by chemical inhibitors, resulted in the artificial induction of the adhE gene under aerobic conditions. Conversely, limiting NAD synthesis, by introducing mutational blocks into the biosynthetic pathway for nicotinic acid, decreased the expression of adhE under anaerobic conditions. This, in turn, was reversed by supplementation with exogenous NAD or nicotinic acid. In merodiploid strains carrying deletion or insertion mutations abolishing the synthesis of AdhE protein, an adhE-lacZ fusion was expressed at nearly 10-fold the level observed in an adhE+ background. Introduction of mutant adhE alleles producing high levels of inactive AdhE protein gave results equivalent to those seen in absence of the AdhE protein. This finding implies that it is the buildup of NADH due to lack of enzyme activity, rather than the absence of the AdhE protein per se, which causes increased induction of the phi(adhE'-lacZ) fusion. Moreover, mutations giving elevated levels of active AdhE protein decreased the induction of the phi(adhE'-lacZ) fusion. This finding suggests that the enzymatic activity of the AdhE protein modulates the level of NADH under anaerobic conditions, thus indirectly regulating its own expression.  相似文献   

9.
The respiratory chain of plant mitochondria differs from that in mammalian mitochondria by containing several rotenone-insensitive NAD(P)H dehydrogenases. Two of these are located on the outer, cytosolic surface of the inner membrane. One is specific for NADH, the other for NADPH. Only the latter is inhibited by diphenyleneiodonium (DPI). Both of these enzymes are normally dependent upon Ca2+ for activity and this constitutes a potentially important mechanism by which the cell can regulate the oxidation of cytosolic NAD(P)H via the concentration of free Ca2+. This and other potential regulatory mechanisms such as the substrate concentration and polyamines are discussed.  相似文献   

10.
S organism ferments ethanol to acetate and H(2) but grows poorly on ethanol unless the partial pressure of H(2) is kept low, as when it is grown in combination with an H(2)-utilizing methanogenic bacterium. The present study shows that S organism contains an alcohol dehydrogenase and a formate dehydrogenase, both of which require nicotinamide adenine dinucleotide (NAD) for activity. Hydrogen is evolved from NADH generated by these activities via a ferredoxin-dependent oxidation of NADH to NAD and H(2). NADH:NADP oxido-reductase activity was also demonstrated. The relationship of these activities to the growth of S organism is discussed.  相似文献   

11.
The oxidation of exogenous NADH by Jerusalem artichoke ( Helianthus tuberosus L.) tuber mitochondria was strongly inhibited at pH 7.2 by EDTA, EGTA and mersalyl and by chlorotetracycline in the presence of Ca2+. This inhibition disappeared at pH 5.5 where about 50% activity was found as compared to controls at pH 7.2. The rate of oxidation of NADPH at pH 5.5 was the same as for NADH but it was inhibited by 50% by both EDTA and mersalyl.
Mitochondria from Arum maculatum spadices oxidised NADH and NADPH with pH optima of 7.2 and 6.5, respectively. In the presence of EDTA the optima shifted to 6.7 and 5.9, respectively, due to an inhibition at higher pH and a lack of inhibition at lower pH. At pH 6.7 NADH oxidation was completely insensitive to both EDTA and mersalyl whereas the oxidation of NADPH was inhibited by more than 50%. The inhibition of NAD(P)H oxidation by chelators at neutral pH was due to the removal of Ca2+ from the membranes in both types of mitochondria. The differences observed in the properties of NADH and NADPH oxidation suggest that two different dehydrogenases are involved. Because of the strong pH-dependence and the changes in chelator-sensitivity in the physiological pH-range 6–8 it is suggested that the properties of NAD(P)H oxidation provide the cell with important means of metabolic regulation.  相似文献   

12.
Metabolic engineering studies have generally focused on manipulating enzyme levels through either the amplification, addition, or deletion of a particular pathway. However, with cofactor-dependent production systems, once the enzyme levels are no longer limiting, cofactor availability and the ratio of the reduced to oxidized form of the cofactor can become limiting. Under these situations, cofactor manipulation may become crucial in order to further increase system productivity. Although it is generally known that cofactors play a major role in the production of different fermentation products, their role has not been thoroughly and systematically studied. However, cofactor manipulations can potentially become a powerful tool for metabolic engineering. Nicotinamide adenine dinucleotide (NAD) functions as a cofactor in over 300 oxidation-reduction reactions and regulates various enzymes and genetic processes. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. This paper investigates a genetic means of manipulating the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase. More specifically, it explores the effect on the metabolic patterns in Escherichia coli under anaerobic and aerobic conditions of substituting the native cofactor-independent formate dehydrogenase (FDH) by and NAD(+)-dependent FDH from Candida boidinii. The over-expression of the NAD(+)-dependent FDH doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed, increased the final cell density, and provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically. Under anaerobic conditions, the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol-to-acetate ratio. Even more interesting is the observation that during aerobic growth, the increased availability of NADH induced a shift to fermentation even in the presence of oxygen by stimulating pathways that are normally inactive under these conditions.  相似文献   

13.
It is generally known that cofactors play a major role in the production of different fermentation products. This paper is part of a systematic study that investigates the potential of cofactor manipulations as a new tool for metabolic engineering. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ and producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. We have previously investigated a genetic means of increasing the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase and have demonstrated that this manipulation provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically (Berríos-Rivera et al., 2002, Metabolic engineering of Escherichia coli: increase of NADH availability by overexpressing an NAD(+)-dependent formate dehydrogenase, Metabolic Eng. 4, 217-229). The current work explores further the effect of substituting the native cofactor-independent formate dehydrogenase (FDH) by an NAD(+)-dependent FDH from Candida boidinii on the NAD(H/+) levels, NADH/NAD+ ratio, metabolic fluxes and carbon-mole yields in Escherichia coli under anaerobic chemostat conditions. Overexpression of the NAD(+)-dependent FDH provoked a significant redistribution of both metabolic fluxes and carbon-mole yields. Under anaerobic chemostat conditions, NADH availability increased from 2 to 3 mol NADH/mol glucose consumed and the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol to acetate ratio and a decrease in the flux to lactate. It was also found that the NADH/NAD+ ratio should not be used as a sole indicator of the oxidation state of the cell. Instead, the metabolic distribution, like the Et/Ac ratio, should also be considered because the turnover of NADH can be fast in an effort to achieve a redox balance.  相似文献   

14.
Based on requirements for acetate or lipoic acid for aerobic (but not anaerobic) growth, Lactococcus lactis subsp. lactis mutants with impaired pyruvate catabolism were isolated following classical mutagenesis. Strains with defects in one or two of the enzymes, pyruvate formate-lyase (PFL), lactate dehydrogenase (LDH) and the pyruvate dehydrogenase complex (PDHC) were obtained. Growth and product formation of these strains were characterized. A PFL-defective strain (requiring acetate for anaerobic growth) displayed a two-fold increase in specific lactate production compared with the corresponding wild-type strain when grown anaerobically. LDH defective strains directed 91-96% of the pyruvate towards alpha-acetolactate, acetoin and diacetyl production when grown aerobically in the presence of acetate and absence of lipoic acid (a similar characteristic was observed in an LDH and PDHC defective strain in the presence of both acetate and lipoic acid) and more than 65% towards formate, acetate and ethanol production under anaerobic conditions. Another strain with defective PFL and LDH was strictly aerobic. However, a variant with strongly enhanced diacetyl reductase activities (NADH/NAD+ dependent diacetyl reductase, acetoin reductase and butanediol dehydrogenase activities) was selected from this strain under anaerobic conditions by supplementing the medium with acetoin. This strain is strictly aerobic, unless supplied with acetoin.  相似文献   

15.
Vanadate-stimulated oxidation of NAD(P)H   总被引:1,自引:0,他引:1  
Vanadate stimulates the oxidation of NAD(P)H by biological membranes because such membranes contain NAD(P)H oxidases which are capable of reducing dioxygen to O2 and because vanadate catalyzes the oxidation of NAD(P)H by O2, by a free radical chain mechanism. Dihydropyridines, such as reduced nicotinamide mononucleotide (NMNH), which are not substrates for membrane-associated NAD(P)H oxidases, are not oxidized by membranes plus vanadate unless NAD(P)H is present to serve as a source of O2. When [NMNH] greatly exceeds [NAD(P)H], in such reaction mixtures, one can observe the oxidation of many molecules of NMNH per NAD(P)H consumed. This reflects the chain length of the free radical chain mechanism. We have discussed the mechanism and significance of this process and have tried to clarify the pertinent but confusing literature.  相似文献   

16.
During starch degradation in intact isolated chloroplasts from Chlamydomonas reinhardtii gas exchange was studied with a mass spectrometer. Oxygen uptake by intact chloroplasts in the dark never exceeded 1.5% of the starch degradation rate [maximum 15 nmol O2 (mg Chl)−1 h−1 consumed. 1 000 nmol glucose (mg Chl)−1h−1 degraded]. Evolution of CO2 under aerobic conditions [9.8–28 nmol (mg Chl)−1 h−1] was stimulated by addition of 0.1–0.5 m M oxaloacetate [393–425 nmol CO2 (mg Chl)−1 h−1]. Pyridoxal phosphate (5 m M ) inhibited starch degradation by more than 80%, but had no effect on O2 uptake. Starch degradation rates and CO2 evolution did not differ under acrobic and anaerobic conditions. Increasing Pi in the reaction medium from 0.5 m M to 5.0 m M stimulated starch degradation by 230 and 260% under aerobic and anaerobic conditions, respectively. A rapid autooxidation of reduced ferredoxin was observed in a reconstituted system consisting of purified Chlamydomonas ferredoxin, purified Chlamydomonas NADP-ferredoxin oxidoreductase (EC 1.6.7.1) and NADPH. Addition of isolated thylakoids from C. reinhardtii did not affect the rate of O2 uptake. Our results clearly indicate the absence of any oxygen requirement during starch degradation in isolated chloroplasts.  相似文献   

17.
The specific growth rate of the ethanol producing bacterium Zymomonas mobilis was 25–40% lower in the presence of oxygen than under anaerobic conditions, provided the cultures were supplied with a low substrate concentration (20 g glucose/l). However, the molar growth yield of these cultures was not influenced by oxygen. With washed cell suspensions, an oxygen consumption could be initiated by the addition of either glucose, fructose, or ethanol. Cell extracts catalyzed the oxidation of NADH with oxygen at a molar ratio of 2:1. Further experiments showed that this NADH oxidase is located in the cell membrane. The specific oxygen consumption rates of cell suspensions correlated with the intracellular NADH oxidizing activities; both levels decreased with increasing concentrations of the fermentation end-product ethanol. The addition of 5 mM NaCN completely inhibited both the intracellular oxygen reduction and also the oxygen consumption of whole cells. Both catalase and superoxide dismutase were present even in anaerobically grown cells. Aeration seemed to have little effect on the level of catalase, but the superoxide dismutase activity was 5-fold higher in cells grown aerobically. Under aerobic conditions considerable amounts of acetaldehyde and acetic acid were formed in addition to the normal fermentation products, ethanol and carbon dioxide.Dedicated to Professor Dr. H. G. Schlegel on the occasion of his 60th birthday  相似文献   

18.
Escherichia coli (E. coli) maintains its total NADH/NAD+ intracellular pool by synthesizing NAD through the de novo pathway and the pyridine nucleotide salvage pathway. The salvage pathway recycles intracellular NAD breakdown products and preformed pyridine compounds from the environment, such as nicotinic acid (NA). The enzyme nicotinic acid phosphoribosyltransferase (NAPRTase; EC 2.4.2.11), encoded by the pncB gene, catalyzes the formation of nicotinate mononucleotide (NAMN), a direct precursor of NAD, from NA. This reaction is believed to be the rate-limiting step in the NAD salvage pathway. The current study investigates the effect of overexpressing the pncB gene from Salmonella typhimurium on the total levels of NAD, the NADH/NAD+ ratio, and the production of different metabolites in E. coli under anaerobic chemostat conditions and anaerobic tube experiments. In addition, this paper studies the effect of combining the overexpression of the pncB gene with an NADH regeneration strategy that increases intracellular NADH availability, as we have previously shown. (The effect of increasing NADH availability on the redistribution of metabolic fluxes in Escherichia coli chemostat cultures, Metabolic Eng. 4, 230-237; Metabolic engineering of Escherichia coli: Increase of NADH availability by overexpressing an NAD(+)-dependent formate dehydrogenase, Metabolic Eng. 4, 217-229.) Overexpression of the pncB gene in chemostat experiments increased the total NAD levels, decreased the NADH/NAD+ ratio, and did not significantly redistribute the metabolic fluxes. However, under anaerobic tube conditions, overexpression of the pncB gene led to a significant shift in the metabolic patterns as evidenced by a decrease in lactate production and an increase as high as two-fold in the ethanol-to-acetate (Et/Ac) ratio. These results suggest that under chemostat conditions the total level of NAD is not limiting and the metabolic rates are fixed by the system at steady state. On the other hand, under transient conditions (such as those in batch cultivation) the increase in the total level of NAD can increase the rate of NADH-dependent pathways (ethanol) and therefore change the final distribution of metabolites. The effect of combining overexpression of the pncB gene with the substitution of the native cofactor-independent formate dehydrogenase (FDH) with an NAD(+)-dependent FDH was also investigated under anaerobic tube conditions. This manipulation produced a metabolic pattern that combines a high Et/Ac ratio similar to that obtained with the new FDH with an intermediate lactate level similar to that obtained with the overexpression of the pncB gene. It was found that addition of the pncB gene to the FDH system does not increase further the production of reduced metabolites because the system for NADH regeneration already reached the maximum theoretical yield of approximately 4 mol NADH/mol of glucose.  相似文献   

19.
Energy conservation in Nitrobacter   总被引:1,自引:0,他引:1  
Abstract The generation of ATP and NADH in total cells of Nitrobacter was measured under aerobic and anaerobic conditions. NADH synthesis was driven by an ATP independent reaction with nitrite or nitric oxide as electron donors. The rate of NADH formation was about 200 times higher, if nitric oxide instead of nitrite served as electron donor. Approximately 2 mol nitric oxide were needed for reduction of 1 mol NAD+. Nitrite caused an end-product inhibition of the nitric oxide induced NADH synthesis. ATP was synthesized by NADH oxidation with oxygen and nitrate as terminal electron acceptors.  相似文献   

20.
Abstract Microbial populations, nitrogen mineralization potentials, and denitrification enzyme activities were examined in two abandoned carbolithic minesoils. Numbers and activities of bacteria and fungi were lower in nonamended than in lime and/or fly ash amended sites. Rates of aerobic NO3 production (3 to 38 μg-N kg−1 h−1) and anaerobic NO3 reduction to N2O (5 to 68 μg-N kg−1 h−1) were measured. Organisms capable of N2O production under anaerobic soil conditions were present in low numbers, and their activity was restricted in part by low soil pH. Nondenitrifying nitrate-reducing bacteria were more diverse and in greater numbers than respiratory denitrifiers and may have been responsible for N2O production in assays measuring denitrification enzyme activity.  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