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.
Two cDNAs encoding NADPH oxidases and constituting the thyroid H(2)O(2) generating system have been cloned. The strategy of cloning was based on the functional similarities between H(2)O(2) generation in leukocytes and the thyroid, according to the hypothesis that one of the components of the thyroid system would belong to the gp91(Phox)/Mox1 gene family and display sequence similarities with gp91(Phox). Screening at low stringency with a gp91(Phox) probe of cDNA libraries from thyroid cells in primary culture yielded two distinct human cDNA clones harboring open reading frames of 1551 (ThOX1) and 1548 amino acids (ThOX2), respectively. The encoded polypeptides display 83% sequence similarity and are clearly related to gp91(Phox) (53 and 47% similarity). The theoretical molecular mass of 177 kDa is close to the apparent molecular mass of 180 kDa of the native corresponding porcine flavoprotein and the protein(s) detected by Western blot in dog and human thyroid. ThOX1 and ThOX2 display sequence similarities of 53% and 61%, respectively, with a predicted protein of Caenorhabditis elegans over their entire length. They show along their first 500 amino acids a similarity of 43% with thyroperoxidase. The corresponding genes of ThOX1 and ThOX2 are closely linked on chromosome 15q15.3. The dog mRNA expression is thyroid-specific and up-regulated by agents activating the cAMP pathway as is the synthesis of the polypeptides they are coding for. In human thyroid the positive regulation by cAMP is less pronounced. The proteins ThOX1 and ThOX2 accumulate at the apical membrane of thyrocytes and are co-localized with thyroperoxidase.  相似文献   

2.
We have recently cloned two thyroid-specific cDNAs encoding new members of the NADPH oxidase family. ThOX1 and ThOX2 proteins are colocalized with thyroperoxidase at the apical membrane of human thyroid cells. In the present study we have determined their subcellular localization and maturation in relation to their enzymatic activity. A majority of ThOX proteins accumulated inside the cell and only a small fraction was expressed at the surface. Western blots demonstrated that ThOX's are glycoproteins of 180,000 and 190,000. When totally deglycosylated the molecular weight of both ThOX1 and ThOX2 drops to 160,000. Ca(2+) stimulates the basal H(2)O(2) generation in PC Cl3 cells at a level corresponding to 20% of the leukocyte H(2)O(2) production stimulated by PMA. Nonthyroid cell lines transfected with ThOX1 and ThOX2 show only a single immunoreactive band in Western blot analysis, corresponding to the protein of 180,000. This "immature" protein remains exclusively intracellular and does not present any enzymatic activity. This is not modified by coexpression of thyroperoxidase and p22(Phox). Transfection of ThOX cDNAs into PLB-XCGD cells does not reconstitute their NADPH oxidase activity. We conclude that (1) the thyroid contains some elements of the leukocyte H(2)O(2)-generating system but not all of them; (2) ThOX's are predominantly or exclusively located inside the cell in thyrocytes or in transfected cells, respectively, and as such they are inactive; (3) ThOX's cannot replace gp91(Phox) in the leukocyte; and (4) the thyroid H(2)O(2)-generating system is analogous to the leukocyte system with regard to ThOX's and gp91(Phox) but very different in other aspects. Additional thyroid-specific components are probably required to get complete protein processing and full enzymatic activity in the thyroid.  相似文献   

3.
4.
An NAD(P)H-dependent H2O2 forming activity has been evidenced in thyroid tissue from patients with Grave's disease. Its biochemical properties were compared to those of the NADPH oxidase previously described in pig thyroid gland. Both were Ca2+-dependent and activated by inorganic phosphate anions in the same range of concentrations. Both are flavoproteins using FAD as cofactor, but the human enzyme was also able to utilize FMN. The apparent Km for NADPH of the human enzyme (100 microM) was 5-10 times higher than that of porcine enzyme. Vm was 3 to 10 times higher in pig (150 nmol x h(-1) x mg(-1)) than in man (14 to 45). Total content in human tissue was 7 to 9% of that in porcine tissue. An unidentified inhibitor has been detected in the 3000 g particulate fraction from most patients, which could account for this apparently low enzyme content. An NADH-dependent H2O2 production has also been observed in porcine and human thyroid tissues. This activity was only partly Ca2+-dependent (man, 50-70%; pig, 80-90%) and presented similar apparent Km values for NADH (man, 100 microM; pig, 200 microM). In pig thyrocytes, the expression of the Ca2+-dependent part of the NADH-oxidase activity was induced by TSH and down-regulated by TGFbeta, as was the NADPH oxidase activity. Furthermore, NADPH and NADH-dependent activities were not additive. We conclude that a single, inducible, NAD(P)H-oxidase can use NADPH or NADH as substrate to catalyse H2O2 formation, and that human and porcine NAD(P)H-oxidases are highly similar. Differences observed could be attributed to minor differences in enzyme structure and/or in membrane microenvironment. The NADH-dependent Ca2+-independent activity observed in human and porcine thyroid fractions could be attributed to a distinct and constitutive enzyme.  相似文献   

5.
Elevated cholesterol levels promote proinflammatory and prothrombogenic responses in venules and impaired endothelium-dependent arteriolar dilation. Although NAD(P)H oxidase-derived superoxide has been implicated in the altered vascular responses to hypercholesterolemia, it remains unclear whether this oxidative pathway mediates the associated arteriolar dysfunction and platelet adhesion in venules. Platelet and leukocyte adhesion in cremasteric postcapillary venules and arteriolar dilation responses to acetylcholine were monitored in wild-type (WT), Cu,Zn-superoxide dismutase transgenic (SOD-TgN), and NAD(P)H oxidase-knockout (gp91(phox-/-)) mice placed on a normal (ND) or high-cholesterol (HC) diet for 2 weeks. HC elicited increased platelet and leukocyte adhesion in WT mice versus ND. Cytosolic subunits of NAD(P)H oxidase (p47phox and p67phox) were expressed in platelets. This was not altered by hypercholesterolemia; however, platelets and leukocytes from HC mice exhibited elevated generation of reactive oxygen species compared to ND mice. Hypercholesterolemia-induced leukocyte recruitment was attenuated in SOD-TgN-HC and gp91(phox-/-)-HC mice. Recruitment of platelets derived from WT-HC mice in venules of SOD-TgN-HC or gp91(phox-/-)-HC recipients was comparable to ND levels. Adhesion of SOD-TgN-HC platelets paralleled the leukocyte response and was attenuated in SOD-TgN-HC recipients, but not in WT-HC recipients. However, gp91(phox-/-)-HC platelets exhibited low levels of adhesion comparable to those of WT-ND in both hypercholesterolemic gp91(phox-/-) and WT recipients. Arteriolar dysfunction was evident in WT-HC mice, compared to WT-ND. Overexpression of SOD or, to a lesser extent, gp91(phox) deficiency restored arteriolar vasorelaxation responses toward WT-ND levels. These findings reveal a novel role for platelet-associated NAD(P)H oxidase in producing the thrombogenic phenotype in hypercholesterolemia and demonstrate that NAD(P)H oxidase-derived superoxide mediates the HC-induced arteriolar dysfunction.  相似文献   

6.
7.
We have previously reported that ANG II stimulation increased superoxide anion (O2-) through the activation of NAD(P)H oxidase and inhibited nitric oxide (NO)-dependent control of myocardial oxygen consumption (MVo2) by scavenging NO. Our objective was to investigate the role of NAD(P)H oxidase, especially the gp91phox subunit, in the NO-dependent control of MVo2. MVo2 in mice with defects in the expression of gp91phox [gp91(phox)(-/-)] was measured with a Clark-type oxygen electrode. Baseline MVo2 was not significantly different between wild-type (WT) and gp91(phox)(-/-) mice. Stimulation of NO production by bradykinin (BK) induced significant decreases in MVo2 in WT mice. BK-induced reduction in MVo2 was enhanced in gp91(phox)(-/-) mice. BK-induced reduction in MVo2 in WT mice was attenuated by 10(-8) mol/l ANG II, which was restored by coincubation with Tiron or apocynin. In contrast to WT mice, BK-induced reduction in MVo2 in gp91(phox)(-/-) mice was not altered by ANG II. There was a decrease in lucigenin (5 x 10(-6) mol/l)-detectable O2- in gp91(phox)(-/-) mice compared with WT mice. ANG II resulted in significant increases in O2- production in WT mice, which was inhibited by coincubation with Tiron or apocynin. However, ANG II had no effect on O2- production in gp91(phox)(-/-) mice. Histological examination showed that the development of abscesses and/or the invasion of inflammatory cells occurred in lungs and livers but not in hearts and kidneys from gp91(phox)(-/-) mice. These results indicate that the gp91(phox) subunit of NAD(P)H oxidase mediates O2- production through the activation of NAD(P)H oxidase and attenuation of NO-dependent control of MVo2 by ANG II.  相似文献   

8.
An NAD(P)H oxidase has been hypothesized to be the main source of reactive oxygen species (ROS) in vessels; however, questions remain about its function and similarity with the neutrophil oxidase. Therefore, vascular superoxide generation was measured by electron paramagnetic resonance spectroscopy using the spin-trap 5,5'-dimethly-pyrroline-N-oxide in aortas from wild-type (WT) and gp91(phox)-deficient mice (gp91(phox)-/-), which do not have a functioning neutrophil NADPH oxidase. There was no significant difference between radical adduct formation by WT or gp91(phox)-/- mouse aortas either at baseline or after stimulation with NADPH or NADH. Also, spin-adduct formation was identical in the 100,000-g pellets obtained from WT and gp91(phox)-/- mouse aortas. SOD mimetics and the flavoenzyme inhibitor diphenyleneiodonium blocked spin-adduct formation from both intact vessels and particulate fractions. Other pharmacological inhibitors of metabolic pathways involved in ROS generation had no effect on this phenomenon. To examine the role of this enzyme in vascular tone control, aortic rings were suspended in organ chambers and preconstricted with phenylephrine to reach half-maximal contraction. Exposure to NADPH elicited a 20% increase in vascular tone, which was decreased by SOD mimetics in a concentration-dependent manner, suggesting that superoxide was responsible for this phenomenon. NADH had no effect on vascular tone. Thus superoxide is generated in the vessel wall by an NAD(P)H-dependent oxidase, which modulates vascular contractile tone. This enzyme is structurally and genetically distinct from the neutrophil NADPH oxidase.  相似文献   

9.
This study was undertaken to reveal the role of NAD(P)H oxidase in increased oxidative stress in islets of Type 2 diabetes. Immunostaining analysis showed that staining intensities of NAD(P)H oxidase components, gp91phox and p22phox, significantly increased in islets of animal models of Type 2 diabetes, OLETF rats (60 weeks of age) and db/db mice (14 weeks of age), compared with age-matched controls, respectively, correlating with increased levels of oxidative stress marker, 8-hydroxy-deoxyguanosine or 4-hydroxy-2-nonenal modified protein. In db/db mice, oral administration of angiotensin II Type 1 receptor antagonist valsartan (5 mg/kg) for 4 weeks significantly attenuated the increased expression of gp91phox and p22phox together with inhibition of oxidative stress and partially restored decreased insulin contents in islets. Angiotensin II-related increased expression of NAD(P)H oxidase may play an important role in increased oxidative stress in islets of Type 2 diabetes. This mechanism may be a novel therapeutic target for preventing beta-cell damage.  相似文献   

10.
Reactive oxygen species (ROS) appear to play an important role in regulating growth and survival of prostate cancer. However, the sources for ROS production in prostate cancer cells have not been determined. We report that ROS are generated by intact American Type Culture Collection DU 145 cells and by their membranes through a mechanism blocked by NAD(P)H oxidase inhibitors. ROS are critical for growth in these cells, because NAD(P)H oxidase inhibitors and antioxidants blocked proliferation. Components of the human phagocyte NAD(P)H oxidase, p22phox and gp91phox, as well as the Ca2+ concentration-responsive gp91phox homolog NOX5 were demonstrated in DU 145 cells by RT-PCR and sequencing. Although the protein product for p22phox was not detectable, both gp91phox and NOX5 were identified throughout the cell by immunostaining and confocal microscopy and NOX5 immunostaining was enhanced in a perinuclear location, corresponding to enhanced ROS production adjacent to the nuclear membrane imaged by 2',7'-dichlorofluorescin diacetate oxidation. The calcium ionophore ionomycin dramatically stimulated ferricytochrome c reduction in cell media, further supporting the importance of NOX5 for ROS production. Antisense oligonucleotides for NOX5 inhibited ROS production and cell proliferation in DU 145 cells. In contrast, antisense oligonucleotides to p22phox or gp91phox did not impair cell growth. Inhibition of ROS generation with antioxidants or NAD(P)H oxidase inhibitors increased apoptosis in cells. These results indicate that ROS generated by the newly described NOX5 oxidase are essential for prostate cancer growth, possibly by providing trophic intracellular oxidant tone that retards programmed cell death. superoxide anion; diphenylene iodonium; p22phox; gp91phox; adenosine 3',5'-cyclic monophosphate response element; caspases  相似文献   

11.
Doxorubicin is a highly effective antineoplastic drug associated with a dose-dependent cardiotoxicity that may result in irreversible cardiomyopathy and heart failure. Gene variants of the superoxide-generating enzyme NAD(P)H oxidase have recently been associated with this phenotype. We investigated the mechanism of this association using lucigenin-enhanced chemiluminescence, spectrophotometry, electrochemical sensor, and electron paramagnetic resonance spectroscopy. Superoxide production was measured in female wild-type and NAD(P)H oxidase-deficient (gp91phox knockout) mice. The magnitude of the increase in superoxide production on the addition of doxorubicin was much higher in hearts of wild-type mice than in enzyme-deficient mice. An increase in superoxide production was observed also on the addition of the NADPH cytochrome P450 reductase. However, doxorubicin reacted with NADPH producing superoxide even in the absence of any enzymatic activity. Taken together, gp91phox-containing NAD(P)H oxidase and NADPH cytochrome P450 reductase can enhance superoxide production caused by the chemical interaction of doxorubicin and NADPH. These findings are in agreement with the recently reported reduced cardiotoxicity following doxorubicin treatment in gp91phox knockout mice and with associations between NAD(P)H oxidase gene variants and sensitivity to doxorubicin.  相似文献   

12.
A novel superoxide-producing NAD(P)H oxidase in kidney   总被引:34,自引:0,他引:34  
During phagocytosis, gp91(phox), the catalytic subunit of the phagocyte NADPH oxidase, becomes activated to produce superoxide, a precursor of microbicidal oxidants. Currently increasing evidence suggests that nonphagocytic cells contain similar superoxide-producing oxidases, which are proposed to play crucial roles in various events such as cell proliferation and oxygen sensing for erythropoiesis. Here we describe the cloning of human cDNA that encodes a novel NAD(P)H oxidase, designated NOX4. The NOX4 protein of 578 amino acids exhibits 39% identity to gp91(phox) with special conservation in membrane-spanning regions and binding sites for heme, FAD, and NAD(P)H, indicative of its function as a superoxide-producing NAD(P)H oxidase. The membrane fraction of kidney-derived human embryonic kidney (HEK) 293 cells, expressing NOX4, exhibits NADH- and NADPH-dependent superoxide-producing activities, both of which are inhibited by diphenylene iodonium, an agent known to block oxygen sensing, and decreased in cells expressing antisense NOX4 mRNA. The human NOX4 gene, comprising 18 exons, is located on chromosome 11q14.2-q21, and its expression is almost exclusively restricted to adult and fetal kidneys. In human renal cortex, high amounts of the NOX4 protein are present in distal tubular cells, which reside near erythropoietin-producing cells. In addition, overexpression of NOX4 in cultured cells leads to increased superoxide production and decreased rate of growth. The present findings thus suggest that the novel NAD(P)H oxidase NOX4 may serve as an oxygen sensor and/or a regulator of cell growth in kidney.  相似文献   

13.
Lee HS  Son SM  Kim YK  Hong KW  Kim CD 《Life sciences》2003,72(24):2719-2730
Reactive oxygen species (ROS) have been implicated in the pathogenesis of vascular dysfunction in diabetes mellitus, and NAD(P)H oxidase is known as the most important source of ROS in the vasculatures. To determine whether NAD(P)H oxidase is a major participant in the critical intermediary signaling events in high glucose (HG, 25 mM)-induced proliferation of vascular smooth muscle cells (VSMC), we investigated in explanted aortic VSMC from rats the role of NAD(P)H oxidase on the HG-related cellular proliferation and superoxide production. VSMC under HG condition had increased proliferative capacity that was inhibited by tiron (1 mM), a cell membrane permeable superoxide scavenger, but not by SOD, which is not permeable to cell membrane. The nitroblue tetrazolium staining in the HG-exposed VSMC was more prominent than that of VSMC under normal glucose (5.5 mM) condition, which was significantly inhibited by DPI (10 microM), an NAD(P)H oxidase inhibitor, but not by inhibitors for other oxidases such as NADH dehydrogenase, xanthine oxidase, and nitric oxide synthase. In the VSMC under HG condition, the enhanced NAD(P)H oxidase activity with increased membrane translocation of Rac1 was observed, but the protein expression of p22phox and gp91phox was not increased. These data suggest that HG-induced changes in VSMC proliferation are related to the intracellular production of superoxide through enhanced activity of NAD(P)H oxidase.  相似文献   

14.
Heme-oxygenase-1 (HO-1), the rate-limiting enzyme of heme degradation, has powerful anti-oxidant properties related to the production of the reactive oxygen species scavenger bilirubin. However, some data suggest that HO-1 could also inhibit the cellular production of reactive oxygen species. Therefore, we investigated whether the anti-oxidant properties of HO-1 could be mediated by modulation of the activity and/or expression of the heme-containing NAD(P)H oxidase, the main source of the superoxide anion (O(2)(-)) in phagocytic cells. Increasing HO-1 expression in RAW 264.7 macrophages effectively decreased NAD(P)H oxidase activity and expression of gp91(phox), its heme-containing catalytic component, because of deficient protein maturation and increased degradation. Loading cells with heme reversed the decrease in O(2)(-) production and gp91(phox) expression induced by HO-1 overexpression. Similar results were obtained in vivo in rat alveolar macrophages after pharmacological modulation of HO-1 expression or activity. These results show that a decrease in heme content due to HO-1 activation limits heme availability for maturation of the gp91(phox) subunit and assembly of the functional NAD(P)H oxidase. This study provides a new mechanism to explain HO-1 anti-oxidant properties.  相似文献   

15.
Primary cytomegalovirus (CMV) infection promotes oxidative stress and reduces nitric oxide (NO) bioavailability in endothelial cells. These events are among the earliest vascular responses to cardiovascular risk factors. We assessed the roles of NAD(P)H oxidase and NO bioavailability in microvascular responses to persistent CMV infection alone or with hypercholesterolemia. Wild-type (WT) or gp91phox (NAD(P)H oxidase subunit) knockout mice received mock inoculum or 3 × 104 PFU murine CMV (mCMV) ip 5 weeks before placement on a normal or high-cholesterol diet (HC) for 4 weeks before assessment of arteriolar function and venular blood cell recruitment using intravital microscopy. Some WT groups received sepiapterin (a precursor of the nitric oxide synthase cofactor tetrahydrobiopterin) or apocynin (NAD(P)H oxidase inhibitor/antioxidant). Endothelium-dependent vasodilation was impaired in mCMV vs mock WT, regardless of diet. This was not affected by sepiapterin, and pharmacological inhibition of nitric oxide synthase reduced dilation similarly in mock and mCMV mice. Apocynin or deficiency of total, but not blood cell or vascular wall only (tested using bone marrow chimeras), gp91phox protected against arteriolar dysfunction. Blood cell recruitment was induced by mCMV–HC. Sepiapterin, but not NAD(P)H oxidase deficiency/apocynin, reduced leukocyte accumulation, whereas platelet adhesion was reduced by sepiapterin, apocynin, or total, platelet-specific, or vascular wall gp91phox deficiency. These data implicate activation of both hematopoietic and vessel wall NAD(P)H oxidase in mCMV-induced arteriolar dysfunction and platelet and vascular NAD(P)H oxidase in the thrombogenic phenotype induced by mCMV–HC. In contrast, findings with sepiapterin suggest that eNOS dysfunction, perhaps uncoupling, mediates venular, but not arteriolar, responses to mCMV–HC, thus indicating that NAD(P)H oxidase and eNOS differentially regulate microvascular responses to mCMV.  相似文献   

16.
Aorta coarctation results in hypertension (HTN) in the arterial tree proximal to stenosis and, as such, provides an ideal model to discern the effects of different levels of blood pressure on the vascular tissue in the same animal. Compelling evidence has emerged supporting the role of oxidative stress as a cause of HTN. However, whether or not HTN (independent of the circulating humoral factors) can cause oxidative stress is less certain. NAD(P)H oxidase isoforms are the main source of reactive oxygen species (ROS) in the vascular tissues. We therefore compared the expressions of NOX-I, gp91phox and the regulatory subunits of the enzyme in the aorta segments residing above and below coarctation in rats with abdominal aorta banding. Rats were studied 4 weeks after aorta banding above the renal arteries or sham operation. Subunits of NAD(P)H oxidase and its NOX-I isoform as well as endothelial NO synthase (eNOS) and nitrotyrosine (footprint of NO oxidation by superoxide) were measured in the aorta segments above and below coarctation. The gp91phox, p47phox, and p67phox subunits of NAD(P)H oxidase, NOX-I isoform, eNOS and nitrotyrosine were markedly increased in the aorta segment above coarctation (hypertensive zone), but were virtually unchanged in the segment below coarctation. Since, excepting blood pressure, all other conditions were constant, the upregulation of NAD(P)H oxidase isoforms and the increased NO oxidation in the aorta segment above, but not below, coarctation prove that HTN, per se, independent of circulating mediators can cause oxidative/nitrosative stress in the arterial wall. These observations suggest that HTN control may represent a specific form of antioxidant therapy for hypertensive disorders.  相似文献   

17.
Hyperoxia increases reactive oxygen species (ROS) production in vascular endothelium; however, the mechanisms involved in ROS generation are not well characterized. We determined the role and regulation of NAD(P)H oxidase in hyperoxia-induced ROS formation in human pulmonary artery endothelial cells (HPAECs). Exposure of HPAECs to hyperoxia for 1, 3, and 12 h increased the generation of superoxide anion, which was blocked by diphenyleneiodonium but not by rotenone or oxypurinol. Furthermore, hyperoxia enhanced NADPH- and NADH-dependent and superoxide dismutase- or diphenyleneiodonium-inhibitable ROS production in HPAECs. Immunohistocytochemistry and Western blotting revealed the presence of gp91, p67 phox, p22 phox, and p47 phox subcomponents of NADPH oxidase in HPAECs. Transfection of HPAECs with p22 phox antisense plasmid inhibited hyperoxia-induced ROS production. Exposure of HPAECs to hyperoxia activated p38 MAPK and ERK, and inhibition of p38 MAPK and MEK1/2 attenuated the hyperoxia-induced ROS generation. These results suggest a role for MAPK in regulating hyperoxia-induced NAD(P)H oxidase activation in HPAECs.  相似文献   

18.
The objective of the present study was to examine the role of the angiotensin II type 1 receptor (AT(1)-R) in the diabetes-aggravated oxidative stress and brain injury observed in a rat model of combined diabetes and focal cerebral ischemia. Diabetes was induced by an injection of streptozotoxin (STZ; 55 mg/kg iv) at 8 wk of age. Two weeks after the induction of diabetes, some animals received continuous subcutaneous infusion of the AT(1)-R antagonist candesartan (0.5 mg.kg(-1).day(-1)) for 14 days. Focal cerebral ischemia, induced by middle cerebral artery occlusion/reperfusion (MCAO), was conducted at 4 wk after STZ injection. Male Sprague-Dawley rats (n = 189) were divided into five groups: normal control, diabetes, MCAO, diabetes + MCAO, and diabetes + MCAO + candesartan. The major observations were that 1) MCAO produced typical cerebral infarction and neurological deficits at 24 h that were accompanied by elevation of NAD(P)H oxidase gp91(phox) and p22(phox) mRNAs, and lipid hydroperoxide production in the ipsilateral hemisphere; 2) diabetes enhanced NAD(P)H oxidase gp91(phox) and p22(phox) mRNA expression, potentiated lipid peroxidation, aggravated neurological deficits, and enlarged cerebral infarction; and 3) candesartan reduced the expression of gp91(phox) and p22(phox), decreased lipid peroxidation, lessened cerebral infarction, and improved the neurological outcome. We conclude that diabetes exaggerates the oxidative stress, NAD(P)H oxidase induction, and brain injury induced by focal cerebral ischemia. The diabetes-aggravated brain injury involves AT(1)-Rs. We have shown for the first time that candesartan reduces brain injury in a combined model of diabetes and cerebral ischemia.  相似文献   

19.
The membrane-integrated protein gp91phox, existing as a heterodimer with p22phox, functions as the catalytic core of the phagocyte NADPH oxidase, which plays a crucial role in host defence. The oxidase, dormant in resting cells, becomes activated to produce superoxide, a precursor of microbicidal oxidants, by interacting with the adaptor proteins p47phox and p67phox as well as the small GTPase Rac. In the past few years, several proteins homologous to gp91phox were discovered as superoxide-producing NAD(P)H oxidases (Nox's) in non-phagocytic cells; however, regulatory mechanisms for the novel oxidases have been largely unknown. Current identification of proteins highly related to p47phox and p67phox, designated Noxol (Nox organizer 1) and Noxal (Nox activator 1), respectively, has shed lights on common and distinct mechanisms underlying activations of Nox family oxidases.  相似文献   

20.
Molecular composition and regulation of the Nox family NAD(P)H oxidases   总被引:12,自引:0,他引:12  
Reactive oxygen species (ROS) are conventionally regarded as inevitable deleterious by-products in aerobic metabolism with a few exceptions such as their significant role in host defense. The phagocyte NADPH oxidase, dormant in resting cells, becomes activated during phagocytosis to deliberately produce superoxide, a precursor of other microbicidal ROS, thereby playing a crucial role in killing pathogens. The catalytic center of this oxidase is the membrane-integrated protein gp91(phox), tightly complexed with p22(phox), and its activation requires the association with p47(phox), p67(phox), and the small GTPase Rac, which normally reside in the cytoplasm. Since recent discovery of non-phagocytic gp91(phox)-related enzymes of the NAD(P)H oxidase (Nox) family--seven homologues identified in humans--deliberate ROS production has been increasingly recognized as important components of various cellular events. Here, we describe a current view on the molecular composition and post-translational regulation of Nox-family oxidases in animals.  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