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 62 毫秒
1.
We studied steps that make up the initial and steady-state phases of nitric oxide (NO) synthesis to understand how activity of bovine endothelial NO synthase (eNOS) is regulated. Stopped-flow analysis of NADPH-dependent flavin reduction showed the rate increased from 0. 13 to 86 s(-1) upon calmodulin binding, but this supported slow heme reduction in the presence of either Arg or N(omega)-hydroxy-l-arginine (0.005 and 0.014 s(-1), respectively, at 10 degrees C). O(2) binding to ferrous eNOS generated a transient ferrous dioxy species (Soret peak at 427 nm) whose formation and decay kinetics indicate it can participate in NO synthesis. The kinetics of heme-NO complex formation were characterized under anaerobic conditions and during the initial phase of NO synthesis. During catalysis heme-NO complex formation required buildup of relatively high solution NO concentrations (>50 nm), which were easily achieved with N(omega)-hydroxy-l-arginine but not with Arg as substrate. Heme-NO complex formation caused eNOS NADPH oxidation and citrulline synthesis to decrease 3-fold and the apparent K(m) for O(2) to increase 6-fold. Our main conclusions are: 1) The slow steady-state rate of NO synthesis by eNOS is primarily because of slow electron transfer from its reductase domain to the heme, rather than heme-NO complex formation or other aspects of catalysis. 2) eNOS forms relatively little heme-NO complex during NO synthesis from Arg, implying NO feedback inhibition has a minimal role. These properties distinguish eNOS from the other NOS isoforms and provide a foundation to better understand its role in physiology and pathology.  相似文献   

2.
A ferrous heme-NO complex builds up in rat neuronal NO synthase during catalysis and lowers its activity. Mutation of a tryptophan located directly below the heme (Trp(409)) to Phe or Tyr causes hyperactive NO synthesis and less heme-NO complex buildup in the steady state (Adak, S., Crooks, C., Wang, Q., Crane, B. R., Tainer, J. A., Getzoff, E. D., and Stuehr, D. J. (1999) J. Biol. Chem. 274, 26907-26911). To understand the mechanism, we used conventional and stopped flow spectroscopy to compare kinetics of heme-NO complex formation, enzyme activity prior to and after complex formation, NO binding affinity, NO complex stability, and its reaction with O(2) in mutants and wild type nNOS. During the initial phase of NO synthesis, heme-NO complex formation was 3 and 5 times slower in W409F and W409Y, and their rates of NADPH oxidation were 50 and 30% that of wild type, probably due to slower heme reduction. NO complex formation slowed NADPH oxidation in the wild type by 7-fold but reduced mutant activities less than 2-fold, giving mutants higher final activities. NO binding kinetics were similar among mutants and wild type, although in ferrous W409Y (and to a lesser extent W409F) the 436-nm NO complex converted to a 417-nm NO complex with time. Oxidation of the ferrous heme-NO complex to ferric enzyme was 7 times faster in Trp(409) mutants than in wild type. Thus, mutant hyperactivity derives from slower formation and faster decay of the heme-NO complex. Together, these minimize partitioning into the NO-bound form.  相似文献   

3.
The heme of neuronal nitric oxide synthase (nNOS) participates in O2 activation but also binds self-generated NO, resulting in reversible feedback inhibition. We utilized mutagenesis to investigate if a conserved tryptophan residue (Trp409), which engages in pi-stacking with the heme and hydrogen bonds to its axial cysteine ligand, helps control catalysis and regulation by NO. Mutants W409F and W409Y were hyperactive regarding NO synthesis without affecting cytochrome c reduction, reductase-independent N-hydroxyarginine oxidation, or Arg and tetrahydrobiopterin binding. In the absence of Arg electron flux through the heme was slower in the W409 mutants than in wild-type. However, less NO complex accumulated during NO synthesis by the mutants. To understand the mechanism, we compared the kinetics of heme-NO complex formation, rate of heme reduction, kcat prior to and after NO complex formation, NO binding affinity, NO complex stability, and its reaction with O2. During the initial phase of NO synthesis, heme-NO complex formation was three and five times slower in W409F and W409Y, which corresponded to a slower heme reduction. NO complex formation inhibited wild-type turnover 7-fold but reduced mutant turnover less than 2-fold, giving mutants higher steady-state activities. NO binding kinetics were similar among mutants and wild type, although mutants also formed a 417 nm ferrous-NO complex. Oxidation of ferrous-NO complex was seven times faster in mutants than in wild type. We conclude that mutant hyperactivity primarily derives from slower heme reduction and faster oxidation of the heme-NO complex by O2. In this way Trp409 mutations minimize NO feedback inhibition by limiting buildup of the ferrous-NO complex during the steady state. Conservation of W409 among NOS suggests that this proximal Trp may regulate NO feedback inhibition and is important for enzyme physiologic function.  相似文献   

4.
After initiating NO synthesis a majority of neuronal NO synthase (nNOS) quickly partitions into a ferrous heme-NO complex. This down-regulates activity and increases enzyme K(m,O(2)). To understand this process, we developed a 10-step kinetic model in which the ferric heme-NO enzyme forms as the immediate product of catalysis, and then partitions between NO dissociation versus reduction to a ferrous heme-NO complex. Rate constants used for the model were derived from recent literature or were determined here. Computer simulations of the model precisely described both pre-steady and steady-state features of nNOS catalysis, including NADPH consumption and NO production, buildup of a heme-NO complex, changes between pre-steady and steady-state rates, and the change in enzyme K(m,O(2)) in the presence or absence of NO synthesis. The model also correctly simulated the catalytic features of nNOS mutants W409F and W409Y, which are hyperactive and display less heme-NO complex formation in the steady state. Model simulations showed how the rate of heme reduction influences several features of nNOS catalysis, including populations of NO-bound versus NO-free enzyme in the steady state and the rate of NO synthesis. The simulation predicts that there is an optimum rate of heme reduction that is close to the measured rate in nNOS. Ratio between NADPH consumption and NO synthesis is also predicted to increase with faster heme reduction. Our kinetic model is an accurate and versatile tool for understanding catalytic behavior and will provide new perspectives on NOS regulation.  相似文献   

5.
Neuronal nitric-oxide synthase (nNOS or NOS I) and endothelial NOS (eNOS or NOS III) differ widely in their reductase and nitric oxide (NO) synthesis activities, electron transfer rates, and propensities to form a heme-NO complex during catalysis. We generated chimeras by swapping eNOS and nNOS oxygenase domains to understand the basis for these differences and to identify structural elements that determine their catalytic behaviors. Swapping oxygenase domains did not alter domain-specific catalytic functions (cytochrome c reduction or H(2)O(2)-supported N(omega)-hydroxy-l-arginine oxidation) but markedly affected steady-state NO synthesis and NADPH oxidation compared with native eNOS and nNOS. Stopped-flow analysis showed that reductase domains either maintained (nNOS) or slightly exceeded (eNOS) their native rates of heme reduction in each chimera. Heme reduction rates were found to correlate with the initial rates of NADPH oxidation and heme-NO complex formation, with the percentage of heme-NO complex attained during the steady state, and with NO synthesis activity. Oxygenase domain identity influenced these parameters to a lesser degree. We conclude: 1) Heme reduction rates in nNOS and eNOS are controlled primarily by their reductase domains and are almost independent of oxygenase domain identity. 2) Heme reduction rate is the dominant parameter controlling the kinetics and extent of heme-NO complex formation in both eNOS and nNOS, and thus it determines to what degree heme-NO complex formation influences their steady-state NO synthesis, whereas oxygenase domains provide minor but important influences. 3) General principles that relate heme reduction rate, heme-NO complex formation, and NO synthesis are not specific for nNOS but apply to eNOS as well.  相似文献   

6.
Nitric oxide (NO) release from nitric oxide synthases (NOSs) depends on the dissociation of a ferric heme-NO product complex (FeIIINO) that forms immediately after NO is made in the heme pocket. The NOS-like enzyme of Bacillus subtilis (bsNOS) has 10-20 fold slower FeIIINO dissociation rate (kd) and NO association rate (kon) compared to mammalian NOS counterparts. We previously showed that an Ile for Val substitution at the opening of the heme pocket in bsNOS contributes to these differences. The complementary mutation in mouse inducible NOS oxygenase domain (Val346Ile) decreased the NO kon and kd by 8 and 3-fold, respectively, compared to wild-type iNOSoxy, and also slowed the reductive processing of the heme-O2 catalytic intermediate. To investigate how these changes affect steady-state catalytic behaviors, we generated and characterized the V346I mutant of full-length inducible NOS (iNOS). The mutant exhibited a 4-5 fold lower NO synthesis activity, an apparent uncoupled NADPH consumption, and formation of a heme-NO complex during catalysis that was no longer sensitive to solution NO scavenging. We found that these altered catalytic behaviors were not due to changes in the heme reduction rate or in the stability of the enzyme heme-O2 intermediate, but instead were due to the slower NO kon and kd and a slower oxidation rate of the enzyme ferrous heme-NO complex. Computer simulations that utilized the measured kinetic values confirmed this interpretation, and revealed that the V346I iNOS has an enhanced NADPH-dependent NO dioxygenase activity that converts almost 1 NO to nitrate for every NO that the enzyme releases into solution. Together, our results highlight the importance of heme pocket geometry in tuning the NO release versus NO dioxygenase activities of iNOS.  相似文献   

7.
We studied catalysis by tetrahydrobiopterin (H4B)-free neuronal nitric-oxide synthase (nNOS) to understand how heme and H4B participate in nitric oxide (NO) synthesis. H4B-free nNOS catalyzed Arg oxidation to N(omega)-hydroxy-l-Arg (NOHA) and citrulline in both NADPH- and H(2)O(2)-driven reactions. Citrulline formation was time- and enzyme concentration-dependent but was uncoupled relative to NADPH oxidation, and generated nitrite and nitrate without forming NO. Similar results were observed when NOHA served as substrate. Steady-state and stopped-flow spectroscopy with the H4B-free enzyme revealed that a ferrous heme-NO complex built up after initiating catalysis in both NADPH- and H(2)O(2)-driven reactions, consistent with formation of nitroxyl as an immediate product. This differed from the H4B-replete enzyme, which formed a ferric heme-NO complex as an immediate product that could then release NO. We make the following conclusions. 1) H4B is not essential for Arg oxidation by nNOS, although it helps couple NADPH oxidation to product formation in both steps of NO synthesis. Thus, the NADPH- or H(2)O(2)-driven reactions form common heme-oxy species that can react with substrate in the presence or absence of H4B. 2) The sole essential role of H4B is to enable nNOS to generate NO instead of nitroxyl. On this basis we propose a new unified model for heme-dependent oxygen activation and H4B function in both steps of NO synthesis.  相似文献   

8.
We have used optical, EPR and M?ssbauer spectroscopies to study the formation of heme-NO complex upon the addition of nitrite to reduced cytochrome cd1 from Thiobacillus denitrificans. The reduced d1 heme binds NO under both alkaline and acidic conditions, but the binding of NO to the reduced c heme was strongly pH-dependent. The M?ssbauer data showed unambiguously that at pH 7.6 the c heme does not complex NO, whereas at pH 5.8 approximately half of the reduced c heme binds NO. This observation was confirmed by EPR studies, which showed that the spin concentration of the heme-NO EPR signal increased from 2 spins/molecule at pH 8.0 to approximately 3 spins/molecule at pH 5.8. Optical absorption study also showed strong pH dependence in the binding of NO to the reduced c heme. We have also analyzed the M?ssbauer spectra of the ferrous d1 heme-NO complex using a spin-Hamiltonian formalism. The magnetic hyperfine coupling tensor was found to be consistent with the unpaired electron residing on a sigma orbital.  相似文献   

9.
Rat neuronal NO synthase (nNOS) contains an Akt-dependent phosphorylation motif in its reductase domain. We mutated a target residue in that site (Ser-1412 to Asp) to mimic phosphorylation and then characterized the mutant using conventional and stopped-flow spectroscopies. Compared with wild-type, S1412D nNOS catalyzed faster cytochrome c and ferricyanide reduction but displayed slower steady-state NO synthesis with greater uncoupling of NADPH oxidation. Paradoxically, the mutant had faster heme reduction, faster heme-NO complex formation, and greater heme-NO complex accumulation at steady state. To understand how these behaviors related to flavin and heme reduction rates, we utilized three soybean calmodulins (CaMs) that supported a range of slower flavin and heme reduction rates in mutant and wild-type nNOS. Reductase activity and two catalytic parameters (speed and amount of heme-NO complex formation) related directly to the speed of flavin and heme reduction. In contrast, steady-state NO synthesis increased, reached a plateau, and then fell at the highest rate of heme reduction that was obtained with S1412D nNOS + CaM. Substituting with soybean CaM slowed heme reduction and increased steady-state NO synthesis by the mutant. We conclude the following. 1) The S1412D mutation speeds electron transfer out of the reductase domain. 2) Faster heme reduction speeds intrinsic NO synthesis but diminishes NO release in the steady state. 3) Heme reduction displays an optimum regarding NO release during steady state. The unique behavior of S1412D nNOS reveals the importance of heme reduction rate in controlling steady-state activity and suggests that nNOS already has a near-optimal rate of heme reduction.  相似文献   

10.
Boggs S  Huang L  Stuehr DJ 《Biochemistry》2000,39(9):2332-2339
To better understand the mechanism of nitric oxide (NO) synthesis, we studied conversion of N-hydroxy-L-arginine (NOHA) or L-arginine (Arg) to citrulline and NO under single-turnover conditions using the oxygenase domain of neuronal nitric oxide synthase (nNOSoxy) and rapid scanning stopped-flow spectroscopy. When anaerobic nNOSoxy saturated with H(4)B and NOHA was provided with 0.5 or 1 electron per heme and then exposed to air at 25 degrees C, it formed 0.5 or 1 mol of citrulline/mol of heme, respectively, indicating that NOHA conversion had 1:1 stoichiometry with respect to electrons added. Identical experiments with Arg produced substoichiometric amounts of NOHA or citrulline even when up to 3 electrons were provided per heme. Transient spectral intermediates were investigated at 10 degrees C. For NOHA, four species were observed in the following sequence: starting ferrous nNOSoxy, a transient ferrous-dioxygen complex, a transient ferric-NO complex, and ferric nNOSoxy. For Arg, transient intermediates other than the ferrous-dioxygen species were not apparent during the reaction. Our results provide a kinetic framework for formation and reactions of the ferrous-dioxygen complex in each step of NO synthesis and establish that (1) the ferrous-dioxy enzyme reacts quantitatively with NOHA but not with Arg and (2) its reaction with NOHA forms 1 NO/heme, which immediately binds to form a ferric heme-NO complex.  相似文献   

11.
Nitric-oxide synthases are flavoheme enzymes that catalyze two sequential monooxygenase reactions to generate nitric oxide (NO) from l-arginine. We investigated a possible redox role for the enzyme-bound cofactor 6R-tetrahydrobiopterin (H4B) in the second reaction of NO synthesis, which is conversion of N-hydroxy-l-arginine (NOHA) to NO plus citrulline. We used stopped-flow spectroscopy and rapid-freeze EPR spectroscopy to follow heme and biopterin transformations during single-turnover NOHA oxidation reactions catalyzed by the oxygenase domain of inducible nitric-oxide synthase (iNOSoxy). Significant biopterin radical (>0.5 per heme) formed during reactions catalyzed by iNOSoxy that contained either H4B or 5-methyl-H4B. Biopterin radical formation was kinetically linked to conversion of a heme-dioxy intermediate to a heme-NO product complex. The biopterin radical then decayed within a 200-300-ms time period just prior to dissociation of NO from a ferric heme-NO product complex. Measures of final biopterin redox status showed that biopterin radical decay occurred via an enzymatic one-electron reduction process that regenerated H4B (or 5MeH4B). These results provide evidence of a dual redox function for biopterin during the NOHA oxidation reaction. The data suggest that H4B first provides an electron to a heme-dioxy intermediate, and then the H4B radical receives an electron from a downstream reaction intermediate to regenerate H4B. The first one-electron transition enables formation of the heme-based oxidant that reacts with NOHA, while the second one-electron transition is linked to formation of a ferric heme-NO product complex that can release NO from the enzyme. These redox roles are novel and expand our understanding of biopterin function in biology.  相似文献   

12.
Wang ZQ  Wei CC  Santolini J  Panda K  Wang Q  Stuehr DJ 《Biochemistry》2005,44(12):4676-4690
Nitric oxide synthases (NOSs) are flavo-heme enzymes that require (6R)-tetrahydrobiopterin (H(4)B) for activity. Our single-catalytic turnover study with the inducible NOS oxygenase domain showed that a conserved Trp that interacts with H(4)B (Trp457 in mouse inducible NOS) regulates the kinetics of electron transfer between H(4)B and an enzyme heme-dioxy intermediate, and this in turn alters the kinetics and extent of Arg hydroxylation [Wang, Z.-Q., et al. (2001) Biochemistry 40, 12819-12825]. To investigate the impact of these effects on NADPH-driven NO synthesis by NOS, we generated and characterized the W457A mutant of inducible NOS and the corresponding W678A and W678F mutants of neuronal NOS. Mutant defects in protein solubility and dimerization were overcome by purifying them in the presence of sufficient Arg and H(4)B, enabling us to study their physical and catalytic profiles. Optical spectra of the ferric, ferrous, heme-dioxy, ferrous-NO, ferric-NO, and ferrous-CO forms of each mutant were similar to that of the wild type. However, the mutants had higher apparent K(m) values for H(4)B and in one mutant for Arg (W457A). They all had lower NO synthesis activities, uncoupled NADPH consumption, and a slower and less prominent buildup of enzyme heme-NO complex during steady-state catalysis. Further analyses showed the mutants had normal or near-normal heme midpoint potential and heme-NO complex reactivity with O(2), but had somewhat slower ferric heme reduction rates and markedly slower reactivities of their heme-dioxy intermediate. We conclude that the conserved Trp (1) has similar roles in two different NOS isozymes and (2) regulates delivery of both electrons required for O(2) activation (i.e., kinetics of ferric heme reduction by the NOS flavoprotein domain and reduction of the heme-dioxy intermediate by H(4)B). However, its regulation of H(4)B electron transfer is most important because this ensures efficient coupling of NADPH oxidation and NO synthesis by NOS.  相似文献   

13.
We compared nitric oxide production and nitrosyl hemoglobin steady state concentrations during the early phases of endotoxemic and hemorrhagic shock of equivalent severity. Sprague-Dawley rats were randomly assigned to (1) sham-operated control, (2) hemorrhage, and (3) intravenous endotoxin. Electron paramagnetic resonance spectroscopy was used to measure NO in the vasculature (binding to hemoglobin) and in the liver (binding to cytochrome P450). Despite similar changes in cardiorespiratory variables and identical microvascular pO(2), nitrosyl hemoglobin concentrations were significantly higher in endotoxemic rats than in rats in hemorrhagic shock, suggesting increased rates of NO production. A substantial venous minus arterial concentration gradient was observed for nitrosyl hemoglobin. This increased in line with the plasma total nitrite + nitrate concentration. Nitrosyl hemoglobin formation is likely to occur predominantly in the venous pool, suggesting that removal of NO from hemoglobin in the presence of oxygen may be faster than previously thought. In the liver, an increase in intracellular heme-NO complexes was detected in endotoxemic rats compared with rats in hemorrhagic shock; this was associated with increased reduction of the mitochondrial respiratory chain and is suggestive of NO inhibition of mitochondrial respiration.  相似文献   

14.
Padma 28 is a mixture of herbs used in traditional Tibetan medicine with anti-inflammatory activities. We investigated the effects of Padma 28 on nitric oxide (NO) production by the inducible nitric oxide synthase (iNOS) in lipopolysaccharide stimulated mouse macrophages (RAW 264.7). Padma 28 (0-900 microg/mL) induced a concentration dependent inhibition of inducible nitric oxide synthesis. iNOS protein expression showed a concentration dependent reduction as revealed by immunoblotting when cells were incubated with increasing amounts of Padma 28. Padma 28 decreased iNOS mRNA levels as shown by RT-PCR. Aqueous extracts from costi amari radix (costus root, the dried root of Saussurea lappa) and the outer cover of myrobalani fructus (the dried fruit of Terminalia chebula), constituents of the complex herb preparation Padma 28, were found to inhibit inducible nitric oxide synthesis by decreasing iNOS protein and iNOS mRNA levels. The inhibition of inducible nitric oxide synthesis might contribute to the anti-inflammatory activities of Padma 28.  相似文献   

15.
Macrophages are commonly cultured at a PO2 of 149 Torr, but tissue macrophages in vivo live in an environment of much lower oxygen tension. Despite the many potential mechanisms for changes in oxygen tension to influence nitric oxide (NO) synthesis, there have been few reports investigating the effect of PO2 on macrophage NO production. With the use of a culture chamber designed to rigorously control oxygen tension, we investigated the effects of culture PO2 on macrophage NO production, inducible nitric oxide synthase (iNOS) activity, iNOS protein, and tumor necrosis factor production. NO production and iNOS activity were linearly related in the range of 39.4 to 677 Torr, but not in the range of 1.03 to 39.4 Torr. Therefore, results obtained in vitro for the high oxygen tensions commonly used in cell culture were quantitatively and qualitatively different from results obtained in cells cultured at the lower oxygen tensions that more accurately reflect the in vivo environment. The influence of oxygen tension on NO production has implications for cell culture methodology and for the relationship between microcirculatory dysfunction and inflammatory responses in rodent models of sepsis.  相似文献   

16.
BACKGROUND: Nitric oxide (NO) is cytostatic for proliferating cells, inhibits microbial growth, and down-regulates the synthesis of specific proteins. Studies were undertaken to determine the mechanism by which NO inhibits total protein synthesis and whether the inhibition correlates with established cytostatic activities of NO. MATERIALS AND METHODS: In in vitro experiments, various cell types were exposed to NO using either donors or expression of inducible NO synthase (iNOS). The capacity of NO to suppress total protein synthesis, measured by incorporation of 35S-methionine into protein, was correlated with the capacity of NO to suppress cell proliferation, viral replication, or iNOS expression. Phosphorylation of eIF-2 alpha was examined as a possible mechanism for the suppressed protein synthesis by NO. RESULTS: Both NO donors and expression of the iNOS suppressed total protein synthesis in L929 cells and A2008 human ovarian tumor cells in parallel with decreased cell proliferation. Suppressed protein synthesis was also shown to correlate with decreased vaccinia virus proliferation in murine peritoneal macrophages in an iNOS-dependent manner. Furthermore, iNOS expression in pancreatic islets or RAW264.7 cells almost completely inhibited total protein synthesis, suggesting that nonspecific inhibition of protein synthesis may be the mechanism by which NO inhibited the synthesis of specific proteins such as insulin or iNOS itself. This possibility was confirmed in RAW264.7 cells where the inhibition of total protein synthesis correlated with the decreased iNOS protein. The decrease in protein levels occurred without changes in iNOS mRNA levels, implicating an inhibition of translation. Mechanistic studies revealed that iNOS expression in RAW264.7 cells resulted in the phosphorylation of eIF-2 alpha and inhibition of the 80S ribosomal complex formation. CONCLUSIONS: These results suggest that NO suppresses protein synthesis by stimulating the phosphorylation of eIF-2 alpha. Furthermore, our observations indicate that nonspecific inhibition of protein synthesis may be a generalized response of cells exposed to high levels of NO and that inhibition of protein synthesis may contribute to many of the described cytostatic actions of NO.  相似文献   

17.
Inducible nitric oxide synthase (iNOS) is an homodimeric enzyme which produces large amounts of nitric oxide (NO) in response to inflammatory stimuli. Several factors affect the synthesis and catalytic activity of iNOS. Particularly, dimerization of NOS monomers is promoted by heme, whereas an intracellular depletion of heme and/or L-arginine considerably decreases NOS resistance to proteolysis. In this study, we found that oxalomalate (OMA, oxalomalic acid, alpha-hydroxy-beta-oxalosuccinic acid), an inhibitor of both aconitase and NADP-dependent isocitrate dehydrogenase, inhibited nitrite production and iNOS protein expression in lipopolysaccharide (LPS)-activated J774 macrophages, without affecting iNOS mRNA content. Furthermore, injection of OMA precursors to LPS-stimulated rats also decreased nitrite production and iNOS expression in isolated peritoneal macrophages. Interestingly, alpha-ketoglutarate or succinyl-CoA administration reversed OMA effect on NO production, thus correlating NO biosynthesis with the anabolic capacity of Krebs cycle. When protein synthesis was blocked by cycloheximide in LPS-activated J774 cells treated with OMA, iNOS protein levels, evaluated by Western blot analysis and (35)S-metabolic labelling, were decreased, suggesting that OMA reduces iNOS biosynthesis and induces an increase in the degradation rate of iNOS protein. Moreover, we showed that OMA inhibits the activity of the iNOS from lung of LPS-treated rats by enzymatic assay. Our results, demonstrating that OMA acts regulating synthesis, catalytic activity and degradation of iNOS, suggest that this compound might have a potential role in reducing the NO overproduction occurring in some pathological conditions.  相似文献   

18.
Nitric oxide (NO), the product of the nitric oxide synthase (NOS) reaction, was previously shown to result in S-nitrosation of the NOS Zn(2+)-tetrathiolate and inactivation of the enzyme. To probe the potential physiological significance of NOS S-nitrosation, we determined the inactivation time scale of the inducible NOS isoform (iNOS) and found it directly correlates with an increase in the level of iNOS S-nitrosation. A kinetic model of NOS inactivation in which arginine is treated as a suicide substrate was developed. In this model, NO synthesized at the heme cofactor is partitioned between release into solution (NO release pathway) and NOS S-nitrosation followed by NOS inactivation (inactivation pathway). Experimentally determined progress curves of NO formation were fit to the model. The NO release pathway was perturbed through addition of the NO traps oxymyoglobin (MbO(2)) and β2 H-NOX, which yielded partition ratios between NO release and inactivation of ~100 at 4 μM MbO(2) and ~22000 at saturating trap concentrations. The results suggest that a portion of the NO synthesized at the heme cofactor reacts with the Zn(2+)-tetrathiolate without being released into solution. Perturbation of the inactivation pathway through addition of the reducing agent GSH or TCEP resulted in a concentration-dependent decrease in the level of iNOS S-nitrosation that directly correlated with protection from iNOS inactivation. iNOS inactivation was most responsive to physiological concentrations of GSH with an apparent K(m) value of 13 mM. NOS turnover that leads to NOS S-nitrosation might be a mechanism for controlling NOS activity, and NOS S-nitrosation could play a role in the physiological generation of nitrosothiols.  相似文献   

19.
20.
Nitric oxide is a free radical gas, NO, of paramount relevance in biology. The enzymes responsible for the synthesis of NO from L-arginine in mammalian tissues are known as nitric oxide synthases (NOS). The inducible NOS (iNOS) is associated with the development of a number of autoimmune diseases. iNOS is induced on monocytes, cells playing a key role in the initiation and progression of the immune response. Induction of the enzyme is effected by proinflammatory cytokines, immunomodulating peptides, and even beta-endorphin through a mechanism involving an increase in cAMP. An excessive production of NO has been implicated in the severe lesions observed in multiple sclerosis (MS). Nitrosation of proteins caused by NO in monocytes may contribute to the formation of new epitopes involved in the autoimmune response. Monocytes/macrophages enhance also their cytotoxic capacity through an increase in NO. iNOS seems to establish a link between neuroendocrine and immune system through beta-endorphin explaining stress-related relapses in MS. One of the causes of demyelination is the lysis of oligodendrocytes by cytotoxic T lymphocytes (CTLs); and T cell response is also known to be modulated by NO.  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