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.
2.
An antisense nitrite reductase (NiR, EC 1.7.7.1) tobacco ( Nicotiana tabacum L.) transformant (clone 271) was used to gain insight into a possible correlation between nitrate reductase (NR, EC 1.6.6.1)-dependent nitrite accumulation and nitric oxide (NO(.)) production, and to assess the regulation of signal transduction in response to stress conditions. Nitrite concentrations of clone 271 leaves were 10-fold, and NO(.) emission rates were 100-fold higher than in wild type leaves. Increased protein tyrosine nitration in clone 271 suggests that high NO(.) production resulted in increased peroxynitrite (ONOO(-)) formation. Tyrosine nitration was also observed in vitro by adding peroxynitrite to leaf extracts. As in mammalian cells, NO(.) and derivatives also increased synthesis of proteins like 14-3-3 and cyclophilins, which are both involved in regulation of activity and stability of enzymes.  相似文献   

3.
Myeloperoxidase (MPO) catalyzes a nitration reaction to form nitrotyrosine in the presence of high nitrite, the metabolite of NO. Human leukocyte was shown to cause phenolic nitration using released MPO as a catalyst in the presence of nitrite. It opposes our previous finding that inhibition of MPO was essential for phenol nitration in human leukocyte study. To clarify the role of MPO, we utilized MPO-deficient human leukocytes and MPO-knockout mice. Even in the absence of exogenously added nitrite, high nitration product was observed in MPO-deficient leukocytes. In liver subjected to ischemia/reperfusion injury, a significantly higher amount of nitrotyrosine was produced in MPO-knockout mice than in normal mice. These results clearly demonstrate that MPO inhibits the accumulation of nitration products in vivo . Further experiments showed that MPO could degrade nitrotyrosine in the presence of glutathione. Thus, MPO-induced degradation of nitration products may cause the underestimation of the nitration product generated in vivo . We conclude that MPO may act predominantly to scavenge nitrotyrosine under physiological nitrite condition, and protect against injurious effect of nitrotyrosine.  相似文献   

4.
Myeloperoxidase (MPO) catalyzes a nitration reaction to form nitrotyrosine in the presence of high nitrite, the metabolite of NO. Human leukocyte was shown to cause phenolic nitration using released MPO as a catalyst in the presence of nitrite. It opposes our previous finding that inhibition of MPO was essential for phenol nitration in human leukocyte study. To clarify the role of MPO, we utilized MPO-deficient human leukocytes and MPO-knockout mice. Even in the absence of exogenously added nitrite, high nitration product was observed in MPO-deficient leukocytes. In liver subjected to ischemia/reperfusion injury, a significantly higher amount of nitrotyrosine was produced in MPO-knockout mice than in normal mice. These results clearly demonstrate that MPO inhibits the accumulation of nitration products in vivo . Further experiments showed that MPO could degrade nitrotyrosine in the presence of glutathione. Thus, MPO-induced degradation of nitration products may cause the underestimation of the nitration product generated in vivo . We conclude that MPO may act predominantly to scavenge nitrotyrosine under physiological nitrite condition, and protect against injurious effect of nitrotyrosine.  相似文献   

5.
Several studies have suggested that exposure to hyperoxia causes lung injury through increased generation of reactive oxygen and nitrogen species. The present study was aimed to investigate the effects of hyperoxia exposure on protein nitration in lungs. Rats were exposed to hyperoxia (>95%) for 48, 60, and 72 h. Histopathological analysis showed a dramatic change in the severity of lung injury in terms of edema and hemorrhage between 48- and 60-h exposure times. Western blot for nitrotyrosine showed that several proteins with molecular masses of 29-66 kDa were nitrated in hyperoxic lung tissues. Immunohistochemical analyses indicate nitrotyrosine staining of alveolar epithelial and interstitial regions. Furthermore, immunoprecipitation followed by Western blot revealed the nitration of surfactant protein A and t1alpha, proteins specific for alveolar epithelial type II and type I cells, respectively. The increased myeloperoxidase (MPO) activity and total nitrite levels in bronchoalveolar lavage and lung tissue homogenates were observed in hyperoxic lungs. Neutrophils and macrophages isolated from the hyperoxia-exposed rats, when cocultured with a rat lung epithelial L2 cell line, caused a significant protein nitration in L2 cells. Inclusion of nitrite further increased the protein nitration. These studies suggest that protein nitration during hyperoxia may be mediated in part by MPO generated from activated phagocytic cells, and such protein modifications may contribute to hyperoxia-mediated lung injury.  相似文献   

6.
The dysfunction and further damage of endothelium play an important role in the development and progression of diabetic vascular complications. Protein tyrosine nitration is involved in endothelial cell injury induced by high glucose. Little is known about protein nitration in human umbilical vein endothelial cells (ECV304) induced by high glucose. In the present article, exposure of ECV304 to 30 mM high glucose (HG30) and 40 mM high glucose (HG40) or hemin–nitrite–H2O2 system for 72 h, the cell injury in ECV304 induced by high glucose and exogenous nitrating agent was studied. After 72 h treatment, it was found that high glucose stimulated ECV304 injury in a dose-dependent manner, including reducing cell viability, increasing malondialdehyde (MDA) content, decreasing glutathione (GSH) content, increasing intracellular reactive oxygen species (ROS), increasing the production of nitric oxygen (NO) (increased nitrite content in cell and nitrate content in medium) and generating protein tyrosine nitration. It was also found that protein tyrosine nitration could induce cell injury further. By comparison the protein tyrosine nitration induced by high glucose condition and extrinsic factors (hemin–nitrite–H2O2 system), it may be speculated that protein is nitrated selectively to generate nitrotyrosine in diabetic vascular complications.  相似文献   

7.
Peroxynitrite, formed in a rapid reaction of nitric oxide (NO) and superoxide anion radical (O(2)), is thought to mediate protein tyrosine nitration in various inflammatory and infectious diseases. However, a recent in vitro study indicated that peroxynitrite exhibits poor nitrating efficiency at biologically relevant steady-state concentrations (Pfeiffer, S., Schmidt, K., and Mayer, B. (2000) J. Biol. Chem. 275, 6346-6352). To investigate the molecular mechanism of protein tyrosine nitration in intact cells, murine RAW 264.7 macrophages were activated with immunological stimuli, causing inducible NO synthase expression (interferon-gamma in combination with either lipopolysaccharide or zymosan A), followed by the determination of protein-bound 3-nitrotyrosine levels and release of potential triggers of nitration (NO, O(2)*, H(2)O(2), peroxynitrite, and nitrite). Levels of 3-nitrotyrosine started to increase at 16-18 h and exhibited a maximum at 20-24 h post-stimulation. Formation of O(2) was maximal at 1-5 h and decreased to base line 5 h after stimulation. Release of NO peaked at approximately 6 and approximately 9 h after stimulation with interferon-gamma/lipopolysaccharide and interferon-gamma/zymosan A, respectively, followed by a rapid decline to base line within the next 4 h. NO formation resulted in accumulation of nitrite, which leveled off at about 50 microm 15 h post-stimulation. Significant release of peroxynitrite was detectable only upon treatment of cytokine-activated cells with phorbol 12-myristate-13-acetate, which led to a 2.2-fold increase in dihydrorhodamine oxidation without significantly increasing the levels of 3-nitrotyrosine. Tyrosine nitration was inhibited by azide and catalase and mimicked by incubation of unstimulated cells with nitrite. Together with the striking discrepancy in the time course of NO/O(2) release versus 3-nitrotyrosine formation, these results suggest that protein tyrosine nitration in activated macrophages is caused by a nitrite-dependent peroxidase reaction rather than peroxynitrite.  相似文献   

8.
It has been reported that mammalian cells incubated with excess nitric oxide (NO) accumulate p53 protein but concomitantly this p53 loses its capacity for binding to its DNA consensus sequence. As nitration of tyrosine residues in various proteins has been shown to inhibit their functions, we examined whether NO nitrates tyrosine residues in p53 protein. MCF-7 cells expressing wild-type p53 were incubated with S-nitrosoglutathione for 4 h and cellular extracts were immunoprecipitated with an anti-p53 antibody. Western blot analyses of immunoprecipitates for p53 or for nitrotyrosine revealed low levels of nitrotyrosine in p53 from untreated cells. Incubation with 2 mM S-nitrosoglutathione induced a significant increase in the nitrotyrosine level in p53 protein compared to nontreated cells. These results suggest that excess NO produced in inflamed tissues could nitrate p53 protein, playing a role in carcinogenesis by impairing functions of this tumor-suppressor protein.  相似文献   

9.
Increased expression of inducible nitric oxide synthase (NOS-2) in inflammatory diseases like uveitis suggests that it contributes to the observed pathological state. The aim of this study was to evaluate corneal expression of NOS-2 and corneal protein nitration in a rat model of uveitis. A single injection of intravitreal lipopolysaccharide was used to induce uveitis. Corneal proteins were separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and visualized by Coomassie blue staining. Expression of NOS-2 and nitrotyrosine (NO(2)Tyr) formation were determined via immunohistochemistry and Western blot analysis. Total nitrate/nitrite levels in the vitreous were measured by spectral analysis via the Griess reagent. Immunohistochemical analysis revealed increased corneal NOS-2 and NO(2)Tyr immunoreactivity in rats with uveitis compared with controls. NOS-2 and NO(2)Tyr immunoreactivity was observed in and around basal cells in the corneal epithelium. Western blot analysis of corneal lysates showed multiple nitrated protein bands in uveitic rats. Spectrophotometric measurement of total nitrate/nitrite levels in the vitreous affirmed significantly increased levels of nitric oxide generation in uveitis (126 +/-2.63 microM/mg protein) compared with controls (65 +/-6.57 microM/mg protein). The presented data suggests that extensive formation of protein nitration and reactive nitrogen species in the cornea contributes to tissue destruction in uveitis. Hence, selective inhibition of NOS-2 may prevent long-term complications and lead to an improvement in the management of uveitis.  相似文献   

10.
Li H  Samouilov A  Liu X  Zweier JL 《Biochemistry》2003,42(4):1150-1159
In addition to nitric oxide (NO) generation from specific NO synthases, NO is also formed during anoxia from nitrite reduction, and xanthine oxidase (XO) catalyzes this process. While in tissues and blood high nitrate levels are present, questions remain regarding whether nitrate is also a source of NO and if XO-mediated nitrate reduction can be an important source of NO in biological systems. To characterize the kinetics, magnitude, and mechanism of XO-mediated nitrate reduction under anaerobic conditions, EPR, chemiluminescence NO-analyzer, and NO-electrode studies were performed. Typical XO reducing substrates, xanthine, NADH, and 2,3-dihydroxybenz-aldehyde, triggered nitrate reduction to nitrite and NO. The rate of nitrite production followed Michaelis-Menten kinetics, while NO generation rates increased linearly following the accumulation of nitrite, suggesting stepwise-reduction of nitrate to nitrite then to NO. The molybdenum-binding XO inhibitor, oxypurinol, inhibited both nitrite and NO production, indicating that nitrate reduction occurs at the molybdenum site. At higher xanthine concentrations, partial inhibition was seen, suggesting formation of a substrate-bound reduced enzyme complex with xanthine blocking the molybdenum site. The pH dependence of nitrite and NO formation indicate that XO-mediated nitrate reduction occurs via an acid-catalyzed mechanism. With conditions occurring during ischemia, myocardial xanthine oxidoreductase and nitrate levels were determined to generate up to 20 microM nitrite within 10-20 min that can be further reduced to NO with rates comparable to those of maximally activated NOS. Thus, XOR catalyzed nitrate reduction to nitrite and NO occurs and can be an important source of NO production in ischemic tissues.  相似文献   

11.
3-nitrotyrosine (NO2-Tyr) is thought to be a specific marker of cell injury during oxidative damage. We have evaluated the role of poly(ADP-ribose)polymerase-1 (PARP-1) in protein nitration after treatment of immortalized fibroblasts parp-1+/+ and parp-1-/- with the alkylating agent 2'-methyl-2'-nitroso-urea (MNU). Both cell lines showed increased iNOS expression following MNU treatment in parallel with a selective induction of tyrosine nitration of different proteins. PARP-1 deficient cells displayed a delayed iNOS accumulation, reduced number of nitrated proteins, and a lower global nitrotyrosine "footprint." We have identified the mitochondrial compartment as the major site of oxidative stress during DNA damage, being MnSOD one of the NO2-Tyr-modified proteins, but not in parp-1-/- cells. These results suggest that NO-derived injury can be modulated by proteins involved in the response to genotoxic damage, such as PARP-1, and may account for the limited oxidative injury in parp-1 knockout mice during carcinogenesis and inflammation.  相似文献   

12.
The recently described L-arginine-dependent nitric oxide (NO) pathway has been proposed to interact synergistically with the TNF pathway in murine macrophage-mediated tumor cytotoxicity in vitro. We have employed an experimental construct in which these two pathways were independently expressed by two different effector cell populations. The TNF-dependent pathway was committed by murine 3T3 cells transfected with the cDNA encoding human pro-TNF. The NO pathway was executed by the murine EMT-6 mammary adenocarcinoma cell line treated with murine rIFN-gamma and LPS. Controls for the TNF pathway committed by the transfectant included lysis of the TNF-sensitive murine L929 cell in coculture, secretion of TNF, and absence of nitrite synthesis. For the NO pathway controls included lysis of the murine P815 mastocytoma cocultured with activated EMT-6 cells that had been pretreated with murine rIFN-gamma and LPS, production of nitrite by this activated effector cell, and an absence of TNF secretion. The target cell panel included L929, EMT-6, P815, and murine B16 melanoma and TU-5 sarcoma cell lines. All targets on this panel were susceptible to lysis by LPS-triggered murine bacillus Calmette-Guérin-activated macrophages. The 3T3 transfectant caused significant lysis of cocultured L929 and TU-5 targets. The EMT-6 effector cell only caused significant lysis of the P815 target. When both effector cells were cocultured with these target cells, lysis of the P815 target was observed to be additive or superadditive; however, for all the other targets, cytotoxicity was comparable with or subadditive compared with that seen with the 3T3 transfectant effector cell alone. Thus, these two pathways do not appear to account for the broad, potent tumoricidal activity observed for activated macrophages in vitro.  相似文献   

13.
The reaction of deoxyhemoglobin with nitric oxide (NO) or nitrite ions (NO 2 (-)) produces iron-nitrosyl-hemoglobin (HbNO) in contrast to the reaction with oxyhemoglobin, which produces methemoglobin and nitrate (NO 3 (-)). HbNO has not been associated with the known bioactivities of NO. We hypothesized that HbNO in erythrocytes could be an important source of bioactive NO/nitrite if its oxidation was coupled to the ascorbic acid (ASC) cycle. Studied by absorption and electron paramagnetic resonance (EPR) spectroscopy, DHA oxidized HbNO to methemoglobin and liberated NO from HbNO as determined by chemiluminescence. Both DHA and ascorbate free radical (AFR), the intermediate between ASC and DHA, enhanced NO oxidation to nitrite, but not nitrate; nor did either oxidize nitrite to nitrate. DHA increased the basal levels of nitrite in erythrocytes, while the reactions of nitrite with hemoglobin are slow. In erythrocytes loaded with HbNO, HbNO disappeared after DHA addition, and the AFR signal was detected by EPR. We suggest that the ASC-AFR-DHA cycle may be coupled to that of HbNO-nitrite and provide a mechanism for the endocrine transport of NO via hemoglobin within erythrocytes, resulting in the production of intracellular nitrite. Additionally, intracellular nitrite and nitrate seem to be largely generated by independent pathways within the erythrocyte. These data provide a physiologically robust mechanism for erythrocytic transport of NO bioactivity allowing for hormone-like properties.  相似文献   

14.
《Free radical research》2013,47(7):883-890
Abstract

This study aimed to determine the effect of haemolysis on plasma oxidation and nitration in sickle cell disease (SCD) patients. Blood was collected from haemoglobin (Hb)A volunteers and homozygous HbSS patients who had not received blood transfusions in the last 3 months. Haemolysis was characterised by low levels of haemoglobin and haptoglobin and high levels of reticulocyte, mean corpuscular volume (MCV), mean corpuscular haemoglobin (MCH), plasma cell-free haemoglobin, bilirubin, total lactate dehydrogenase (LDH) and dominance of LDH-1 isoenzyme. Plasma 8-isoprostane, protein carbonyl and nitrotyrosine levels were measured to evaluate oxidised lipids, oxidised and nitrated proteins, respectively. Plasma nitrite–nitrate levels were also determined to assess nitric oxide (NO) production in both SCD patients and controls. Markers of haemolysis were significantly evident in SCD patients compared to controls. Plasma 8-isoprostane, protein carbonyl and nitrotyrosine levels were markedly elevated in SCD patients compared to controls. Linear regression analysis revealed a significant inverse correlation between haemoglobin and reticulocyte counts and a significant positive correlation of plasma cell-free haemoglobin with protein carbonyl and nitrotyrosine levels. The obtained data shows that increased haemolysis in SCD increases plasma protein oxidation and nitration.  相似文献   

15.
Nitric oxide (NO) can stimulate dendritic cells to a more activated state. However, nitric oxide and peroxynitrites production by dendritic cells has been usually associated with pathological situations such as autoimmunity or inflammatory diseases. This study was designed to determine if dendritic cells obtained from healthy volunteers produce nitric oxide and peroxynitrites, which results in protein nitration. The expression of arginase II, but not arginase I, isoform was detected in monocytes and dendritic cells. There was higher inducible nitric oxide synthase (iNOS) protein expression and lower arginase activity both in immature and mature dendritic cells, compared to monocytes. This caused nitric oxide production, and maturation of dendritic cells which provoked a significative increase of nitrites and nitrates compared to immature dendritic cells. There was also peroxynitrites synthesis during monocyte differentiation as shown by the nitration of proteins. Immunoblot revealed a pattern of nitrated proteins in cell extracts obtained from monocytes and dendritic cells, however there were bands that appeared only in human dendritic cells, in particular an intense 90 kDa band. Nitric oxide production and nitrotyrosine formation could affect the antigen presentation and modify the immune response.  相似文献   

16.
We examined the relative contributory roles of extracellular vs. intracellular L-arginine (ARG) toward cellular activation of endothelial nitric oxide synthase (eNOS) in human endothelial cells. EA.hy926 human endothelial cells were incubated with different concentrations of (15)N(4)-ARG, ARG, or L-arginine ethyl ester (ARG-EE) for 2h. To modulate ARG transport, siRNA for ARG transporter (CAT-1) vs. sham siRNA were transfected into cells. ARG transport activity was assessed by cellular fluxes of ARG, (15)N(4)-ARG, dimethylarginines, and L-citrulline by an LC-MS/MS assay. eNOS activity was determined by nitrite/nitrate accumulation, either via a fluorometric assay or by(15)N-nitrite or estimated (15)N(3)-citrulline concentrations when (15)N(4)-ARG was used to challenge the cells. We found that ARG-EE incubation increased cellular ARG concentration but no increase in nitrite/nitrate was observed, while ARG incubation increased both cellular ARG concentration and nitrite accumulation. Cellular nitrite/nitrate production did not correlate with cellular total ARG concentration. Reduced (15)N(4)-ARG cellular uptake in CAT-1 siRNA transfected cells vs. control was accompanied by reduced eNOS activity, as determined by (15)N-nitrite, total nitrite and (15)N(3)-citrulline formation. Our data suggest that extracellular ARG, not intracellular ARG, is the major determinant of NO production in endothelial cells. It is likely that once transported inside the cell, ARG can no longer gain access to the membrane-bound eNOS. These observations indicate that the "L-arginine paradox" should not consider intracellular ARG concentration as a reference point.  相似文献   

17.
Endoplasmic reticulum (ER) stress and excessive nitric oxide production via the induction of inducible nitric oxide synthase (NOS2) have been implicated in the pathogenesis of ocular diseases characterized by retinal degeneration. Previous studies have revealed the sphingomyelinase/ceramide pathway in the regulation of NOS2 induction. Thus, the objective of this study was to determine the activity of the sphingomyelinase/ceramide pathway, assess nitric oxide production, and examine apoptosis in human retinal pigment epithelial (RPE) cells undergoing ER stress. Sphingomyelinase (SMase) activity; nuclear factor κB (NF-κB) activation; NOS2, nitrite/nitrate, and nitrotyrosine levels; and apoptosis were determined in cultured human RPE cell lines subjected to ER stress via exposure to tunicamycin. Induction of ER stress was confirmed by increased intracellular levels of ER stress markers including phosphorylated PKR-like ER kinase, C/EBP-homologous protein, and 78-kDa glucose-regulated protein. ER stress increased nuclear translocation of NF-κB, NOS2 expression, nitrite/nitrate levels, and nitrotyrosine formation and caused apoptosis in RPE cell lines. Inhibition of neutral SMase (N-SMase) activity via GW 4869 treatment caused a significant reduction in nuclear translocation of NF-κB, NOS2 expression, nitrite/nitrate levels, nitrotyrosine formation, and apoptosis in ER-stressed RPE cells. In conclusion, N-SMase inhibition reduced nitrative stress and apoptosis in RPE cells undergoing ER stress. Obtained data suggest that NOS2 can be regulated by N-SMase in RPE cells experiencing ER stress.  相似文献   

18.
The role of nitric oxide (NO) in graft-versus-tumor-effect (GVT) was evaluated in the present study. GVT was induced by intravenous injection of C57BL/6J (H-2b) mouse splenocytes to {C57BL/6J (H-2b) x BALB/c (H-2d)} F1 mice bearing Meth A (H-2d) ascites tumors. Induction of GVT increased nitrite production and expression of inducible NO synthase by ascites cells. The increased nitrite production was inhibited by NG-monomethyl-L-arginine (MLA). Experiments employing immunomagnetic depletion of Mac-1+ cells from ascites indicated that macrophages were a major cellular source of the nitrite production. Interferon-gamma levels were increased in both serum and ascites fluid during GVT. Induction of GVT prolonged survival of ascites-bearing mice, and increased urinary nitrate excretion. MLA administration inhibited GVT-induced increase in urinary nitrate excretion, and further prolonged GVT-induced increase in survival. These results indicate that NO synthesis is induced in tumors during GVT, and the NO acts as an inhibitor of GVT.  相似文献   

19.
Peroxynitrite-mediated tyrosine nitration catalyzed by superoxide dismutase.   总被引:69,自引:0,他引:69  
Peroxynitrite (ONOO-), the reaction product of superoxide (O2-) and nitric oxide (NO), may be a major cytotoxic agent produced during inflammation, sepsis, and ischemia/reperfusion. Bovine Cu,Zn superoxide dismutase reacted with peroxynitrite to form a stable yellow protein-bound adduct identified as nitrotyrosine. The uv-visible spectrum of the peroxynitrite-modified superoxide dismutase was highly pH dependent, exhibiting a peak at 438 nm at alkaline pH that shifts to 356 nm at acidic pH. An equivalent uv-visible spectrum was obtained by Cu,Zn superoxide dismutase treated with tetranitromethane. The Raman spectrum of authentic nitrotyrosine was contained in the spectrum of peroxynitrite-modified Cu,Zn superoxide dismutase. The reaction was specific for peroxynitrite because no significant amounts of nitrotyrosine were formed with nitric oxide (NO), nitrogen dioxide (NO2), nitrite (NO2-), or nitrate (NO3-). Removal of the copper from the Cu,Zn superoxide dismutase prevented formation of nitrotyrosine by peroxynitrite. The mechanism appears to involve peroxynitrite initially reacting with the active site copper to form an intermediate with the reactivity of nitronium ion (NO2+), which then nitrates tyrosine on a second molecule of superoxide dismutase. In the absence of exogenous phenolics, the rate of nitration of tyrosine followed second-order kinetics with respect to Cu,Zn superoxide dismutase concentration, proceeding at a rate of 1.0 +/- 0.1 M-1.s-1. Peroxynitrite-mediated nitration of tyrosine was also observed with the Mn and Fe superoxide dismutases as well as other copper-containing proteins.  相似文献   

20.
It was recently reported that chloramphenicol inhibits existing denitrification enzyme activity in sediments and carbon-starved cultures of "Pseudomonas denitrificans." Therefore, we studied the effect of chloramphenicol on denitrification by Flexibacter canadensis and "P. denitrificans." Production of N(inf2)O from nitrate by F. canadensis cells decreased as the concentration of chloramphenicol was increased, and 10.0 mM chloramphenicol completely inhibited N(inf2)O production. "P. denitrificans" was less sensitive to chloramphenicol, and production of N(inf2)O from nitrate was inhibited by only about 50% even in the presence of 10.0 mM chloramphenicol. These results suggested that inhibition of denitrification enzyme activity depended on the concentration of chloramphenicol. Increasing the concentration of chloramphenicol decreased the rate of production of nitrite from nitrate by F. canadensis cells, and the concentration of chloramphenicol which resulted in 50% inhibition of production of nitrite from nitrate was 2.5 mM. In contrast, the rates of production of nitrite from nitrate by intact cells and cell extracts of "P. denitrificans" were inhibited by only 58 and 54%, respectively, at a chloramphenicol concentration of 10.0 mM. Chloramphenicol caused accumulation of NO from nitrite but not from nitrate and inhibited NO consumption in F. canadensis; however, it had neither effect in "P. denitrificans." Chloramphenicol did not affect N(inf2)O consumption by these organisms. We concluded that chloramphenicol inhibits denitrification at the level of nitrate reduction and, in F. canadensis, also at the level of NO reduction.  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