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 0 毫秒
1.
Neuronal nitric-oxide synthase (nNOS) is activated by the Ca(2+)-dependent binding of calmodulin (CaM) to a characteristic polypeptide linker connecting the oxygenase and reductase domains. Calmodulin binding also activates the reductase domain of the enzyme, increasing the rate of reduction of external electron acceptors such as cytochrome c. Several unusual structural features appear to control this activation mechanism, including an autoinhibitory loop, a C-terminal extension, and kinase-dependent phosphorylation sites. Pre-steady state reduction and oxidation time courses for the nNOS reductase domain indicate that CaM binding triggers NADP(+) release, which may exert control over steady-state turnover. In addition, the second order rate constant for cytochrome c reduction in the absence of CaM was found to be highly dependent on the presence of NADPH. It appears that NADPH induces a conformational change in the nNOS reductase domain, restricting access to the FMN by external electron acceptors. CaM binding reverses this effect, causing a 30-fold increase in the second order rate constant. The results show a startling interplay between the two ligands, which both exert control over the conformation of the domain to influence its electron transfer properties. In the full-length enzyme, NADPH binding will probably close the conformational lock in vivo, preventing electron transfer to the oxygenase domain and the resultant stimulation of nitric oxide synthesis.  相似文献   

2.
3.
The mechanisms by which nitric-oxide synthases (NOSs) bind and activate oxygen at their P450-type heme active site in order to synthesize nitric oxide from the substrate L-arginine are mostly unknown. To obtain information concerning the structure and properties of the first oxygenated intermediate of the enzymatic cycle, we have used a rapid continuous flow mixer and resonance Raman spectroscopy to generate and identify the ferrous dioxygen complex of the oxygenase domain of nNOS (Fe(2+)O(2) nNOSoxy). We detect a line at 1135 cm(-1) in the resonance Raman spectrum of the intermediate formed from 0.6 to 3.0 ms after the rapid mixing of the ferrous enzyme with oxygen that is shifted to 1068 cm(-1) with (18)O(2). This line is assigned as the O-O stretching mode (nu(O-O)) of the oxygenated complex of nNOSoxy. Rapid mixing experiments performed with nNOSoxy saturated with L-arginine or N(omega)-hydroxy-L-arginine, in the presence or absence of (6R)-5,6, 7,8-tetrahydro-L-biopterin, reveal that the nu(O-O) line is insensitive to the presence of the substrate and the pterin. The optical spectrum of this ferrous dioxygen species, with a Soret band wavelength maximum at 430 nm, confirms the identification of the previously reported oxygenated complexes generated by stopped flow techniques.  相似文献   

4.
One-electron reduction of chromate by NADPH-dependent glutathione reductase   总被引:2,自引:0,他引:2  
Electron spin resonance (ESR) measurements provide evidence for the formation of Cr(V) intermediates in the enzymatic reduction of Cr(VI) by glutathione reductase (GSSG-R) in the presence of NADPH, indicating an initial single-electron transfer step in the reduction mechanism. Depending on the pH, at least two different Cr(V) species are generated which are relatively long-lived. In addition, we have detected the hydroxyl (.OH) radical formation during the GSSG-R catalyzed reduction of Cr(VI) by spin trapping, employing 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and alpha-(4-pyridyl-1-oxide)-N-tert-butylnitrone (4-POBN) as spin traps. Superoxide dismutase (SOD) causes only a minor effect on the .OH radical and Cr(V) formation, indicating that the O2- is not significantly involved in the reaction mechanism. Catalase enhances the Cr(V) formation and substantially inhibits the .OH radical formation, indicating the involvement of hydrogen peroxide (H2O2) in the reaction mechanism. Addition of H2O2 suppresses Cr(V) and enhances the .OH radical formation. Measurements involving N-ethylmaleimide show that the Cr(V) species, produced enzymatically by the reduction of Cr(VI) by GSSG-R, react with H2O2 to generate .OH radicals, which might participate in the initiation of Cr(VI) carcinogenicity.  相似文献   

5.
The nNOS reductase domain is homologous to cytochrome P450 reductase, which contains two conserved clusters of acidic residues in its FMN module that play varied roles in its electron transfer reactions. To study the role of nNOS reductase domain cluster 1 acidic residues, we mutated two conserved acidic (Asp(918) and Glu(919)) and one conserved aromatic residue (Phe(892)), and investigated the effect of each mutation on flavin binding, conformational change, electron transfer reactions, calmodulin regulation, and catalytic activities. Each mutation destabilized FMN binding without significantly affecting other aspects including substrate, cofactor or calmodulin binding, or catalytic activities upon FMN reconstitution, indicating the mutational effect was restricted to the FMN module. Characterization of the FMN-depleted mutants showed that bound FMN was essential for reduction of the nNOS heme or cytochrome c, but not for ferricyanide or dichlorophenolindolphenol, and established that the electron transfer path in nNOS is NADPH to FAD to FMN to heme. Steady-state and stopped-flow kinetic analysis revealed a novel role for bound FMN in suppressing FAD reduction by NADPH. The suppression could be relieved either by FMN removal or calmodulin binding. Calmodulin binding induced a conformational change that was restricted to the FMN module. This increased the rate of FMN reduction and triggered electron transfer to the heme. We propose that the FMN module of nNOS is the key positive or negative regulator of electron transfer at all points in nNOS. This distinguishes nNOS from other related flavoproteins, and helps explain the mechanism of calmodulin regulation.  相似文献   

6.
The nitric-oxide synthases (NOSs) are comprised of an oxygenase domain and a reductase domain bisected by a calmodulin (CaM) binding region. The NOS reductase domains share approximately 60% sequence similarity with the cytochrome P450 oxidoreductase (CYPOR), which transfers electrons to microsomal cytochromes P450. The crystal structure of the neuronal NOS (nNOS) connecting/FAD binding subdomains reveals that the structure of the nNOS-connecting subdomain diverges from that of CYPOR, implying different alignments of the flavins in the two enzymes. We created a series of chimeric enzymes between nNOS and CYPOR in which the FMN binding and the connecting/FAD binding subdomains are swapped. A chimera consisting of the nNOS heme domain and FMN binding subdomain and the CYPOR FAD binding subdomain catalyzed significantly increased rates of cytochrome c reduction in the absence of CaM and of NO synthesis in its presence. Cytochrome c reduction by this chimera was inhibited by CaM. Other chimeras consisting of the nNOS heme domain, the CYPOR FMN binding subdomain, and the nNOS FAD binding subdomain with or without the tail region also catalyzed cytochrome c reduction, were not modulated by CaM, and could not transfer electrons into the heme domain. A chimera consisting of the heme domain of nNOS and the reductase domain of CYPOR reduced cytochrome c and ferricyanide at rates 2-fold higher than that of native CYPOR, suggesting that the presence of the heme domain affected electron transfer through the reductase domain. These data demonstrate that the FMN subdomain of CYPOR cannot effectively substitute for that of nNOS, whereas the FAD subdomains are interchangeable. The differences among these chimeras most likely result from alterations in the alignment of the flavins within each enzyme construct.  相似文献   

7.
Mechanism of superoxide generation by neuronal nitric-oxide synthase   总被引:8,自引:0,他引:8  
Neuronal nitric-oxide synthase (NOS I) in the absence of L-arginine has previously been shown to generate superoxide (O-2) (Pou, S., Pou, W. S., Bredt, D. S., Snyder, S. H., and Rosen, G. M. (1992) J. Biol. Chem. 267, 24173-24176). In the presence of L-arginine, NOS I produces nitric oxide (NO.). Yet the competition between O2 and L-arginine for electrons, and by implication formation of O-2, has until recently remained undefined. Herein, we investigated this relationship, observing O-2 generation even at saturating levels of L-arginine. Of interest was the finding that the frequently used NOS inhibitor NG-monomethyl L-arginine enhanced O-2 production in the presence of L-arginine because this antagonist attenuated NO. formation. Whereas diphenyliodonium chloride inhibited O-2, blockers of heme such as NaCN, 1-phenylimidazole, and imidazole likewise prevented the formation of O-2 at concentrations that inhibited NO. formation from L-arginine. Taken together these data demonstrate that NOS I generates O-2 and the formation of this free radical occurs at the heme domain.  相似文献   

8.
Phosphorylation of neuronal nitric-oxide synthase (nNOS) by Ca2+/calmodulin (CaM)-dependent protein kinases (CaM kinases) including CaM kinase Ialpha (CaM-K Ialpha), CaM kinase IIalpha (CaM-K IIalpha), and CaM kinase IV (CaM-K IV), was studied. It was found that purified recombinant nNOS was phosphorylated by CaM-K Ialpha, CaM-K IIalpha, and CaM-K IV at Ser847 in vitro. Replacement of Ser847 with Ala (S847A) prevented phosphorylation by CaM kinases. Phosphorylated recombinant wild-type nNOS at Ser847 (approximately 0.5 mol of phosphate incorporation into nNOS) exhibited a 30% decrease of Vmax with little change of both the Km for L-arginine and Kact for CaM relative to unphosphorylated enzyme. The activity of mutant S847D was decreased to a level 50-60% as much as the wild-type enzyme. The decreased NOS enzyme activity of phosphorylated nNOS at Ser847 and mutant S847D was partially due to suppression of CaM binding, but not to impairment of dimer formation which is thought to be essential for enzyme activation. Inactive nNOS lacking CaM-binding ability was generated by mutation of Lys732-Lys-Leu to Asp732-Asp-Glu (Watanabe, Y., Hu, Y., and Hidaka, H. (1997) FEBS Lett. 403, 75-78). It was phosphorylated by CaM kinases, as was the wild-type enzyme, indicating that CaM-nNOS binding was not required for the phosphorylation reaction. We developed antibody NP847, which specifically recognize nNOS in its phosphorylated state at Ser847. Using the antibody NP847, we obtained evidence that nNOS is phosphorylated at Ser847 in rat brain. Thus, our results suggest that CaM kinase-induced phosphorylation of nNOS at Ser847 alters the activity control of this enzyme.  相似文献   

9.
Neurons that express neuronal nitric-oxide synthase (nNOS) are resistant to NO-induced neurotoxicity; however, the mechanism by which these neurons are protected is not clear. To identify proteins possibly involved in this process, we performed affinity chromatography with the nNOS PDZ domain, a N-terminal motif that mediates protein interactions. Using this method to fractionate soluble tissue extracts, we identified the muscle isoform of phosphofructokinase (PFK-M) as a protein that binds to nNOS both in brain and skeletal muscle. PFK-M interacts with the PDZ domain of nNOS, and nNOS-PFK-M binding can be competed by peptides that bind to the PDZ domain of nNOS. We found that nNOS is significantly associated with PFK-M in skeletal muscle because nNOS can be immunodepleted from cytosolic skeletal muscle extracts using an antibody directed against PFK-M. In brain, nNOS and PFK-M are both enriched in synaptosomes, and specifically, in the synaptic vesicle fraction, where they can interact. At the cellular level, PFK-M is enriched in neurons that express nNOS protein. As fructose-1, 6-bisphosphate, the product of PFK activity, is neuroprotective, the interaction of nNOS and PFK may contribute to neuroprotection of nNOS positive cells.  相似文献   

10.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  相似文献   

11.
It is established that aminoguanidine (AG) is a metabolism-based inactivator of the three major isoforms of nitric-oxide synthase. AG is thought to be of potential use in diseases, such as diabetes, where pathological overproduction of NO is implicated. We show here that during the inactivation of neuronal nitric-oxide synthase (nNOS) by AG that the prosthetic heme is altered, in part, to dissociable and protein-bound adducts. The protein-bound heme adduct is the result of cross-linking of the heme to residues in the oxygenase domain of nNOS. The dissociable heme product is unstable and reverts back to heme upon isolation. The alteration of the heme is concomitant with the loss in the ability to form the ferrous-CO complex of nNOS and accounts for at least two-thirds of the activity loss. Studies with [(14)C]AG indicate that alteration of the protein, in part on the reductase domain of nNOS, also occurs but at low levels. Thus, heme alteration appears to be the major cause of nNOS inactivation. The elucidation of the mechanism of inactivation of nNOS will likely lead to a better understanding of the in vivo effects of NOS inhibitors such as AG.  相似文献   

12.
Nitric-oxide synthase (NOS) is a fusion protein composed of an oxygenase domain with a heme-active site and a reductase domain with an NADPH binding site and requires Ca(2+)/calmodulin (CaM) for NO formation activity. We studied NO formation activity in reconstituted systems consisting of the isolated oxygenase and reductase domains of neuronal NOS with and without the CaM binding site. Reductase domains with 33-amino acid C-terminal truncations were also examined. These were shown to have faster cytochrome c reduction rates in the absence of CaM. N(G)-hydroxy-l-Arg, an intermediate in the physiological NO synthesis reaction, was found to be a viable substrate. Turnover rates for N(G)-hydroxy-l-Arg in the absence of Ca(2+)/CaM in most of the reconstituted systems were 2.3-3.1 min(-1). Surprisingly, the NO formation activities with CaM binding sites on either reductase or oxygenase domains were decreased dramatically on addition of Ca(2+)/CaM. However, NADPH oxidation and cytochrome c reduction rates were increased by the same procedure. Activation of the reductase domains by CaM addition or by C-terminal deletion failed to increase the rate of NO synthesis. Therefore, both mechanisms appear to be less important than the domain-domain interaction, which is controlled by CaM binding in wild-type neuronal NOS, but not in the reconstituted systems.  相似文献   

13.
Cu(2+) and Zn(2+) inhibit all of the NADPH-dependent reactions catalyzed by neuronal nitric-oxide synthase (nNOS) including ferricytochrome c reduction, NADPH oxidation, and citrulline formation. Cu(2+) and Zn(2+) also inhibit ferricytochrome c reduction by the independent reductase domain. Zn(2+) affects all activities of the full-length nNOS and the reductase domain to the same extent (estimated IC(50) values from 9 to 31 microm), suggesting Zn(2+) occupation of a single site in the reductase domain. Citrulline formation and NADPH oxidation by the full-length nNOS and ferricytochrome c reduction by the reductase domain are affected similarly by Cu(2+), with estimated IC(50) values ranging from 6 to 33 microm. However, Cu(2+) inhibits ferricytochrome c reduction by the full-length nNOS 2 orders of magnitude more potently, with an estimated IC(50) value of 0.12 microm. These data suggest the possibility that Cu(2+) may interact with nNOS at two sites, one composed exclusively of the reductase domain (which is perhaps also involved in Zn(2+)-mediated inhibition), and another that includes components of both domains. Occupation of the second (higher affinity) site could then promote the selective inhibition of ferricytochrome c reduction in full-length nNOS. Neither the inhibition by Cu(2+) nor that by Zn(2+) is dependent on calmodulin.  相似文献   

14.
The reduction of flavin in hepatic NADH-cytochrome b5 reductase by the hydrated electron (eaq-) was investigated by pulse radiolysis. The eaq- reduced the flavin of NADH-cytochrome b5 reductase to form the red semiquinone between pH 5 and 9. The spectrum of the red semiquinone differs from that of enzyme reduced by dithionite in the presence of NAD+. After the first phase of the reduction, conversion of the red to blue semiquinone was observed at acidic pH. Resulting products are the blue (neutral) or red (anionic) semiquinone or a mixture of the two forms. The pK value for this flavin radical was approximately 6.3. Subsequently, the semiquinone form reacted by dismutation to form the oxidized and the fully reduced forms of the enzyme with a rate constant of 1 x 10(3) M-1 s-1 at pH 7.1. In the presence of NAD+, eaq- reacted with NAD+ to yield NAD(.). Subsequently, NAD. transferred an electron to NAD+-bound oxidized enzyme to form the blue and red semiquinone or mixture of the two forms of the enzyme, where pK value of this flavin radical was approximately 6.3. The blue semiquinone obtained at acidic pH was found to convert to the red semiquinone with a first order rate constant of 90 s-1, where the rates were not affected by pH or the concentration of NAD+. The final product is NAD+-bound red semiquinone of the enzyme.  相似文献   

15.
At glutamatergic synapses, the scaffolding protein PSD95 links the neuronal isoform of nitric-oxide synthase (nNOS) to the N-methyl-d-aspartate (NMDA) receptor. Phosphorylation of nNOS at serine 847 (Ser(847)) by the calcium-calmodulin protein kinase II (CaMKII) inhibits nNOS activity, possibly by blocking the binding of Ca(2+)-CaM. Here we show that the NMDA mediates a novel bidirectional regulation of Ser(847) phosphorylation. nNOS phosphorylated at Ser(847) colocalizes with the NMDA receptor at spines of cultured hippocampal neurons. Treatment of neurons with 5 microm glutamate stimulated CaMKII phosphorylation of nNOS at Ser(847), whereas excitotoxic concentrations of glutamate, 100 and 500 microm, induced Ser(847)-PO(4) dephosphorylation by protein phosphatase 1. Strong NMDA receptor stimulation was likely to activate nNOS under these conditions because protein nitration to form nitrotyrosine, a marker of nNOS activity, correlated in individual neurons with Ser(847)-PO(4) dephosphorylation. Of particular note, stimulation with low glutamate that increased phosphorylation of nNOS at Ser(847) could be reversed by subsequent high glutamate treatment which induced dephosphorylation. The reversibility of NMDA receptor-induced phosphorylation at Ser(847) by different doses of glutamate suggests two mechanisms with opposite effects: 1). a time-dependent negative feedback induced by physiological concentrations of glutamate that limits nNOS activation and precludes the overproduction of NO; and 2). a pathological stimulation by high concentrations of glutamate that leads to unregulated nNOS activation and production of toxic levels of NO. These mechanisms may share pathways, respectively, with NMDA receptor-induced forms of synaptic plasticity and excitotoxicity.  相似文献   

16.
We report here the unexpected finding that recombinant or hepatic microsomal NADPH-cytochrome P450 reductase catalyzes the oxidative deformylation of a model xenobiotic aldehyde, 2-phenylpropionaldehyde, to the n-1 alcohol, 1-phenylethanol, in the absence of cytochrome P450. The flavoprotein and NADPH are absolute requirements, and the reaction displays a dependence on time and on NADPH and reductase concentration. Not surprisingly, the hydrophobic tail of the flavoprotein is not required for catalytic competence. The reductase domain of neuronal nitric oxide synthase is about 30% more active than P450 reductase, and neither flavoprotein catalyzes conversion of the aldehyde to the carboxylic acid, by far the predominant metabolite with P450s in a reconstituted system. Reductase-catalyzed deformylation is unaffected by metal ion chelators and oxygen radical scavengers, but is strongly inhibited by catalase, and the catalase-mediated inhibition is prevented by azide. These results, together with observed parallel increases in 1-phenylethanol and H(2)O(2) formation as a function of NADPH concentration, are evidence that free H(2)O(2) is rate-limiting in aldehyde deformylation by the flavoprotein reductases. This contrasts sharply with the P450-catalyzed reaction, which is brought about by iron-bound peroxide that is inaccessible to catalase.  相似文献   

17.
Neuronal nitric oxide synthase (nNOS) is composed of an oxygenase domain that binds heme, (6R)-tetrahydrobiopterin, and Arg, coupled to a reductase domain that binds FAD, FMN, and NADPH. Activity requires dimeric interaction between two oxygenase domains and calmodulin binding between the reductase and oxygenase domains, which triggers electron transfer between flavin and heme groups. We constructed four different nNOS heterodimers to determine the path of calmodulin-induced electron transfer in a nNOS dimer. A predominantly monomeric mutant of rat nNOS (G671A) and its Arg binding mutant (G671A/E592A) were used as full-length subunits, along with oxygenase domain partners that either did or did not contain the E592A mutation. The E592A mutation prevented Arg binding to the oxygenase domain in which it was present. It also prevented NO synthesis when it was located in the oxygenase domain adjacent to the full-length subunit. However, it had no effect when present in the full-length subunit (i.e. the subunit containing the reductase domain). The active heterodimer (G671A/E592A full-length subunit plus wild type oxygenase domain subunit) showed remarkable similarity with wild type homodimeric nNOS in its catalytic responses to five different forms and chimeras of calmodulin. This reveals an active involvement of calmodulin in supporting transelectron transfer between flavin and heme groups on adjacent subunits in nNOS. In summary, we propose that calmodulin functions to properly align adjacent reductase and the oxygenase domains in a nNOS dimer for electron transfer between them, leading to NO synthesis by the heme.  相似文献   

18.
Ubiquitination of neuronal nitric-oxide synthase in vitro and in vivo   总被引:2,自引:0,他引:2  
It is established that suicide inactivation of neuronal nitric-oxide synthase (nNOS) with guanidine compounds, or inhibition of the hsp90-based chaperone system with geldanamycin, leads to the enhanced proteolytic degradation of nNOS. This regulated proteolysis is mediated, in part, by the proteasome. We show here with the use of human embryonic kidney 293 cells transfected with nNOS that inhibition of the proteasome with lactacystin leads to the accumulation of immunodetectable higher molecular mass forms of nNOS. Some of these higher molecular mass forms were immunoprecipitated by an anti-ubiquitin antibody, indicating that they are nNOS-polyubiquitin conjugates. Moreover, the predominant nNOS-ubiquitin conjugate detected in human embryonic kidney 293 cells, as well as in rat brain cytosol, migrates on SDS-polyacrylamide gels with a mobility near that for the native monomer of nNOS and likely represents a conjugate containing a few or perhaps one ubiquitin. Studies in vitro with the use of (125)I-ubiquitin and reticulocyte extracts could mimic this ubiquitination reaction, which was dependent on ATP. The heme-deficient monomeric form of nNOS is preferentially ubiquitinated over that of the heme-sufficient functionally active homodimer. Thus, we have shown for the first time that ubiquitination of nNOS occurs and is likely involved in the regulated proteolytic removal of non-functional enzyme.  相似文献   

19.
Guanabenz, a metabolism-based irreversible inactivator of neuronal nitric-oxide synthase (nNOS) in vitro, causes the loss of immunodetectable nNOS in vivo. This process is selective in that the slowly reversible inhibitor N(G)-nitro-L-arginine did not decrease the levels of nNOS in vivo. To better understand the mechanism for the loss of nNOS protein in vivo, we have investigated the effects of guanabenz and N(G)-nitro-L-arginine in HEK 293 cells stably transfected with the enzyme. We show here that guanabenz, but not N(G)-nitro-L-arginine, caused the inactivation and loss of nNOS protein in the HEK 293 cells. In studies with cycloheximide or in pulse-chase experiments with [(35)S]methionine, we demonstrate that the loss of nNOS was due in large part to enhanced proteolysis of the protein with the half-life decreasing by one-half from 20 to 10 h. Other metabolism-based irreversible inactivators to nNOS, N(G)-methyl-L-arginine, and N(5)-(1-iminoethyl)-L-ornithine, but not the reversible inhibitor 7-nitroindazole (7-NI), caused a similar decrease in the half-life of nNOS. Proteasomal inhibitors, lactacystin, Cbz-leucine-leucine-leucinal, and N-acetyl-leucine-leucine-norleucinal, but not the lysosomal protease inhibitor leupeptin, were found to effectively inhibit the proteolytic degradation of nNOS. Thus we have shown for the first time that the irreversible inactivators of nNOS, perhaps through covalent alteration of the enzyme, enhance the proteolytic turnover of the enzyme by a mechanism involving the proteasome.  相似文献   

20.
It is established that neuronal nitric-oxide synthase (nNOS) is ubiquitylated and proteasomally degraded. The proteasomal degradation of nNOS is enhanced by suicide inactivation of nNOS or by the inhibition of hsp90, which is a chaperone found in a native complex with nNOS. In the current study, we have examined whether CHIP, a chaperone-dependent E3 ubiquitin-protein isopeptide ligase that is known to ubiquitylate other hsp90-chaperoned proteins, could act as an ubiquitin ligase for nNOS. We found with the use of HEK293T or COS-7 cells and transient transfection methods that CHIP overexpression causes a decrease in immunodetectable levels of nNOS. The extent of the loss of nNOS is dependent on the amount of CHIP cDNA used for transfection. Lactacystin (10 microM), a selective proteasome inhibitor, attenuates the loss of nNOS in part by causing the nNOS to be found in a detergent-insoluble form. Immunoprecipitation of the nNOS and subsequent Western blotting with an anti-ubiquitin IgG shows an increase in nNOS-ubiquitin conjugates because of CHIP. Moreover, incubation of nNOS with a purified system containing an E1 ubiquitin-activating enzyme, an E2 ubiquitin carrier protein conjugating enzyme (UbcH5a), CHIP, glutathione S-transferase-tagged ubiquitin, and an ATP-generating system leads to the ubiquitylation of nNOS. The addition of purified hsp70 and hsp40 to this in vitro system greatly enhances the amount of nNOS-ubiquitin conjugates, suggesting that CHIP is an E3 ligase for nNOS whose action is facilitated by (and possibly requires) its interaction with nNOS-bound hsp70.  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