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.
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   

2.
Homodimer formation activates all nitric-oxide synthases (NOSs). It involves the interaction between two oxygenase domains (NOSoxy) that each bind heme and (6R)-tetrahydrobiopterin (H4B) and catalyze NO synthesis from L-Arg. Here we compared three NOSoxy isozymes regarding dimer strength, interface composition, and the ability of L-Arg and H4B to stabilize the dimer, promote its formation, and protect it from proteolysis. Urea dissociation studies indicated that the relative dimer strengths were NOSIIIoxy > NOSIoxy > NOSIIoxy (endothelial NOSoxy (eNOSoxy) > neuronal NOSOXY (nNOSoxy) > inducible NOSoxy (iNOSoxy)). Dimer strengths of the full-length NOSs had the same rank order as judged by their urea-induced loss of NO synthesis activity. NOSoxy dimers containing L-Arg plus H4B exhibited the greatest resistance to urea-induced dissociation followed by those containing either molecule and then by those containing neither. Analysis of crystallographic structures of eNOSoxy and iNOSoxy dimers showed more intersubunit contacts and buried surface area in the dimer interface of eNOSoxy than iNOSoxy, thus revealing a potential basis for their different stabilities. L-Arg plus H4B promoted dimerization of urea-generated iNOSoxy and nNOSoxy monomers, which otherwise was minimal in their absence, and also protected both dimers against trypsin proteolysis. In these respects, L-Arg alone was more effective than H4B alone for nNOSoxy, whereas for iNOSoxy the converse was true. The eNOSoxy dimer was insensitive to proteolysis under all conditions. Our results indicate that the three NOS isozymes, despite their general structural similarity, differ markedly in their strengths, interfaces, and in how L-Arg and H4B influence their formation and stability. These distinguishing features may provide a basis for selective control and likely help to regulate each NOS in its particular biologic milieu.  相似文献   

3.
The effect of temperature (between 5.0 and 45.0 degrees C) on the catalytic activity of nitric oxide synthases types I, II, and III (NOS-I, NOS-II, and NOS-III, respectively) has been investigated, at pH 7.5. The value of V(max) for NOS-I activity increases from 1.8 x 10(1) pmol min(-1) mg(-1), at 5.0 degrees C, to 1.8 x 10(2) pmol min(-1) mg(-1), at 45.0 degrees C; on the other hand, the value of K(m) (=4.0 x 10(-6) M) is temperature independent. Again, the value of V(max) for NOS-II activity increases from 8.0 pmol min(-1) mg(-1), at 7.0 degrees C, to 5.4 x 10(1) pmol min(-1) mg(-1), at 40.0 degrees C, the value of K(m) (=1.8 x 10(-5) M) being unaffected by temperature. Temperature exerts the same effect on NOS-I and NOS-II activity, as shown by the same values of DeltaH(V(max)) (=4.2 x 10(1) kJ mol(-1)), DeltaH(K(m)) (=0 kJ mol(-1)), and DeltaH((V(max))(/K(m))()) (=4.2 x 10(1) kJ mol(-1)). On the contrary, the value of K(m) for NOS-III activity decreases from 3.8 x 10(-5) M, at 10.0 degrees C, to 1.6 x 10(-5) M, at 40.0 degrees C, the value of V(max) (=6.8 x 10(1) pmol min(-1) mg(-1)) being temperature independent. Present results indicate that temperature influences directly NOS-I and NOS-II activity independently of the substrate concentration, the values of K(m) being temperature independent. However, when l-arginine level is higher than 2 x 10(-4) M, as observed under in vivo conditions, NOS-III activity is essentially unaffected by temperature, the substrate concentration exceeding the value of K(m). As a whole, although further studies in vivo are needed, these observations seem to have potential physiopathologic implications.  相似文献   

4.
Bacterial nitric-oxide synthase proteins (NOSs) from certain Streptomyces strains have been shown to participate in biosynthetic nitration of tryptophanyl moieties in vivo (Kers, J. A., Wach, M. J., Krasnoff, S. B., Cameron, K. D., Widom, J., Bukhaid, R. A., Gibson, D. M., and Crane, B. R., and Loria, R. (2004) Nature 429, 79-82). We report that the complex between Deinococcus radiodurans NOS (deiNOS) and an unusual tryptophanyl-tRNA synthetase (TrpRS II) catalyzes the regioselective nitration of tryptophan (Trp) at the 4-position. Unlike non-enzymatic Trp nitration, and similar reactions catalyzed by globins and peroxidases, deiNOS only produces the otherwise unfavorable 4-nitro-Trp isomer. Although deiNOS alone will catalyze 4-nitro-Trp production, yields are significantly enhanced by TrpRS II and ATP. 4-Nitro-Trp formation exhibits saturation behavior with Trp (but not tyrosine) and is completely inhibited by the addition of the mammalian NOS cofactor (6R)-5,6,7,8-tetrahydro-l-biopterin (H(4)B). Trp stimulates deiNOS oxidation of substrate l-arginine (Arg) to the same degree as H(4)B. These observations are consistent with a mechanism where Trp or a derivative thereof binds in the NOS pterin site, participates in Arg oxidation, and becomes nitrated at the 4-position.  相似文献   

5.
We previously reported the existence of a special auto-regulation property of neuronal nitric-oxide synthase (NOS) based on NO near-geminate combination and partial trapping of neuronal NOS (nNOS) through a futile regenerating pathway. On this basis, we developed a kinetic simulation model that was proven to predict nNOS catalytic specificities and mutations effects (Santolini, J., Adak, S., Curran, C. M., and Stuehr, D. J. (2001) J. Biol. Chem. 276, 1233-1243; Adak, S., Santolini, J., Tikunova, S., Wang, Q., Johnson, J. D., and Stuehr, D. J. (2001) J. Biol. Chem. 276, 1244-1252). Here we show that the same model simulates and explains the distinct catalytic behaviors of inducible and endothelial NOS (iNOS and eNOS). Their marked differences were linked to variations in three basic parameters (rates of ferric heme reduction, ferric heme.NO dissociation, and ferrous heme.NO oxidation) that together control partitioning between futile and productive pathways and their relative rates. We also incorporated feedback inhibition into the kinetic model to account for potential rebinding of accumulated solution NO. The model accurately simulated the different relative impacts of both NOS.NO interactions (near-geminate combination of NO versus rebinding of solution NO) on catalytic behavior of each NOS isoform, including their speed and extent of heme.NO complex accumulation, K(m) for O(2), and propensity to transform NO into a higher oxide. Thus, individual catalytic behavior of any NOS can be understood through a single unified kinetic model. Because the model defines how different settings of individual kinetic parameters control regulation by two distinct NOS.NO interactions, it sheds light on mechanisms, structural features, and scope of NOS regulation and its physiologic impact.  相似文献   

6.
Neuronal nitric-oxide synthase (nNOS) is composed of a heme oxygenase domain and a flavin-bound reductase domain. Ca(2+)/calmodulin (CaM) is essential for interdomain electron transfer during catalysis, whereas the role of the catalytically important cofactor, tetrahydrobiopterin (H4B) remains elusive. The product NO appears to bind to the heme and works as a feedback inhibitor. The present study shows that the Fe(3+)-NO complex is reduced to the Fe(2+)-NO complex by NADPH in the presence of both l-Arg and H4B even in the absence of Ca(2+)/CaM. The complex could not be fully reduced in the absence of H4B under any circumstances. However, dihydrobiopterin and N(G)-hydroxy-l-Arg could be substituted for H4B and l-Arg, respectively. No direct correlation could be found between redox potentials of the nNOS heme and the observed reduction of the Fe(3+)-NO complex. Thus, our data indicate the importance of the pterin binding to the active site structure during the reduction of the NO-heme complex by NADPH during catalytic turnover.  相似文献   

7.
P S Brzovi?  K Ngo  M F Dunn 《Biochemistry》1992,31(15):3831-3839
Tryptophan synthase from enteric bacteria is an alpha 2 beta 2 bienzyme complex that catalyzes the final two reactions in the biosynthesis of L-tryptophan (L-Trp) from 3-indole-D-glycerol 3'-phosphate (IGP) and L-serine (L-Ser). The bienzyme complex exhibits reciprocal ligand-mediated allosteric interactions between the heterologous subunits [Houben, K., & Dunn, M. F. (1990) Biochemistry 29, 2421-2429], but the relationship between allostery and catalysis had not been completely defined. We have utilized rapid-scanning stopped-flow (RSSF) UV-visible spectroscopy to study the relationship between allostery and catalysis in the alpha beta-reaction catalyzed by the bienzyme complex from Salmonella typhimurium. The pre-steady-state spectral changes that occur when L-Ser and IGP are mixed simultaneously with the alpha 2 beta 2 complex show that IGP binding to the alpha-site accelerates the formation of alpha-aminoacrylate [E(A-A)] from L-Ser at the beta-site. Through the use of L-Ser analogues, we show herein that the formation of the E(A-A) intermediate is the chemical signal which triggers the conformational transition that activates the alpha-subunit. beta-subunit ligands, such as L-Trp, that react to form covalent intermediates at the beta-site, but are incapable of E(A-A) formation, do not stimulate the activity of the alpha-subunit. Titration experiments show that the affinity of G3P and GP at the alpha-site is dependent upon the nature of the chemical intermediate present at the beta-active site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
9.
Nitric oxide (NO) is synthesized from L-Arg in the P450-type heme active site of nitric-oxide synthase (NOS). The internal axial ligand of the heme, Cys415, may hydrogen-bond to the side chain of the conserved Arg418 residue in neuronal NOS (nNOS). To understand the role of Arg418, we generated the nNOS mutants, Arg418Ala and Arg418Leu. NO formation activities with the mutants using both L-Arg and NHA as substrates were less than 0.1 nmol/min/nmol heme, in contrast to rates of 34-35 nmol/min/nmol heme with the wild-type enzyme. The heme reduction rate of the mutants was very slow, less than 10(-2) min(-1), in contrast with that (more than 10 min(-1)) of the wild type. The backbone amide group of Arg418 interacts with the Cys415 thiolate through van der Waals contact, whereas the carbonyl oxygen of Cys415 and the guanidino N(epsilon) atom of Arg418 form a tight hydrogen bond. The results suggest that Arg418 is critical in preserving the heme proximal structure and thus, is indirectly involved in both catalysis and electron transfer from the reductase domain to the heme.  相似文献   

10.
Recently, we characterized a novel endothelial nitric-oxide synthase (eNOS)-interacting protein, NOSTRIN (for eNOS-trafficking inducer), which decreases eNOS activity upon overexpression and induces translocation of eNOS away from the plasma membrane. Here, we show that NOSTRIN directly binds to caveolin-1, a well-established inhibitor of eNOS. Because this interaction occurs between the N terminus of caveolin (positions 1-61) and the central domain of NOSTRIN (positions 323-434), it allows for independent binding of each of the two proteins to eNOS. Consistently, we were able to demonstrate the existence of a ternary complex of NOSTRIN, eNOS, and caveolin-1 in Chinese hamster ovary (CHO)-eNOS cells. In human umbilical vein endothelial cells (HUVECs), the ternary complex assembles at the plasma membrane upon confluence or thrombin stimulation. In CHO-eNOS cells, NOSTRIN-mediated translocation of eNOS involves caveolin in a process most likely representing caveolar trafficking. Accordingly, trafficking of NOSTRIN/eNOS/caveolin is affected by altering the state of actin filaments or cholesterol levels in the plasma membrane. During caveolar trafficking, NOSTRIN functions as an adaptor to recruit mediators such as dynamin-2 essential for membrane fission. We propose that a ternary complex between NOSTRIN, caveolin-1, and eNOS mediates translocation of eNOS, with important implications for the activity and availability of eNOS in the cell.  相似文献   

11.
A strong interaction between iron(III) and calf thymus DNA at pH 7.4 was demonstrated in the present study by separation of the complex by column chromatography and by the slow kinetics of iron(III) removal from DNA by disodium-1,2-dihydroxybenzene-3,5-disulfonate (Tiron). An equilibrium constant of 2.1 x 10(14) was calculated by measurements of bound iron(III) by flame atomic absorption spectroscopy and assuming a one iron to two nucleotide stoichiometry. Graphic analysis of the interaction however, indicated that DNA has two binding sites for iron(III) characterized by a stoichiometry of one iron to 12 nucleotides and one iron to 2 nucleotides, and association constants of 4.8 x 10(12) and 2.3 x 10(11), respectively. The DNA-iron(III) complex isolated by column chromatography was shown to catalyze the oxidation of both 2-phenylethylhydrazine and methylhydrazine by spin-trapping experiments with alpha-(4-pyridyl 1-oxide)-N-tert-butylnitrone (POBN). By contrast, oxidation of 1,2-dimethylhydrazine was not catalyzed. Catalysis of 2-phenylethylhydrazine oxidation was confirmed by oxygen consumption studies. The results suggest that iron chelated to DNA may be significant in DNA damage induced by oxidizable chemicals.  相似文献   

12.
The present study shows that rat liver and brain mitochondrial nitric oxide synthase (mtNOS) are functionally associated with mitochondrial respiratory chain complex I. When complex I is activated, mtNOS exerts high activity and generates nitric oxide, whereas inactivation of complex I leads mtNOS to abandon its NOS activity. Functional association of mtNOS with complex I is potentially important in regulating mtNOS activity and mitochondrial functions.  相似文献   

13.
14.
The main role of the plasma membrane Ca2+/calmodulin-dependent ATPase (PMCA) is in the removal of Ca2+ from the cytosol. Recently, we and others have suggested a new function for PMCA as a modulator of signal transduction pathways. This paper shows the physical interaction between PMCA (isoforms 1 and 4) and alpha-1 syntrophin and proposes a ternary complex of interaction between endogenous PMCA, alpha-1 syntrophin, and NOS-1 in cardiac cells. We have identified that the linker region between the pleckstrin homology 2 (PH2) and the syntrophin unique (SU) domains, corresponding to amino acids 399-447 of alpha-1 syntrophin, is crucial for interaction with PMCA1 and -4. The PH2 and the SU domains alone failed to interact with PMCA. The functionality of the interaction was demonstrated by investigating the inhibition of neuronal nitric-oxide synthase-1 (NOS-1); PMCA is a negative regulator of NOS-1-dependent NO production, and overexpression of alpha-1 syntrophin and PMCA4 resulted in strongly increased inhibition of NO production. Analysis of the expression levels of alpha-1 syntrophin protein in the heart, skeletal muscle, brain, uterus, kidney, or liver of PMCA4-/- mice, did not reveal any differences when compared with those found in the same tissues of wild-type mice. These results suggest that PMCA4 is tethered to the syntrophin complex as a regulator of NOS-1, but its absence does not cause collapse of the complex, contrary to what has been reported for other proteins within the complex, such as dystrophin. In conclusion, the present data demonstrate for the first time the localization of PMCA1b and -4b to the syntrophin.dystrophin complex in the heart and provide a specific molecular mechanism of interaction as well as functionality.  相似文献   

15.
The mitochondrial oxidative phosphorylation (OXPHOS) system consists of four electron transport chain (ETC) complexes (CI–CIV) and the FoF1-ATP synthase (CV), which sustain ATP generation via chemiosmotic coupling. The latter requires an inward-directed proton-motive force (PMF) across the mitochondrial inner membrane (MIM) consisting of a proton (ΔpH) and electrical charge (Δψ) gradient. CI actively participates in sustaining these gradients via trans-MIM proton pumping. Enigmatically, at the cellular level genetic or inhibitor-induced CI dysfunction has been associated with Δψ depolarization or hyperpolarization. The cellular mechanism of the latter is still incompletely understood. Here we demonstrate that chronic (24 h) CI inhibition in HEK293 cells induces a proton-based Δψ hyperpolarization in HEK293 cells without triggering reverse-mode action of CV or the adenine nucleotide translocase (ANT). Hyperpolarization was associated with low levels of CII-driven O2 consumption and prevented by co-inhibition of CII, CIII or CIV activity. In contrast, chronic CIII inhibition triggered CV reverse-mode action and induced Δψ depolarization. CI- and CIII-inhibition similarly reduced free matrix ATP levels and increased the cell's dependence on extracellular glucose to maintain cytosolic free ATP. Our findings support a model in which Δψ hyperpolarization in CI-inhibited cells results from low activity of CII, CIII and CIV, combined with reduced forward action of CV and ANT.  相似文献   

16.
The nitric-oxide synthases (NOSs) make nitric oxide and citrulline from l-arginine. How the bound cofactor (6R)-tetrahydrobiopterin (H4B) participates in Arg hydroxylation is a topic of interest. We demonstrated previously that H4B radical formation in the inducible NOS oxygenase domain (iNOSoxy) is kinetically coupled to the disappearance of a heme-dioxy intermediate and to Arg hydroxylation. Here we report single turnover studies that determine and compare the kinetics of these transitions in Arg hydroxylation reactions catalyzed by the oxygenase domains of endothelial and neuronal NOSs (eNOSoxy and nNOSoxy). There was a buildup of a heme-dioxy intermediate in eNOSoxy and nNOSoxy followed by a monophasic transition to ferric enzyme during the reaction. The rate of heme-dioxy decay matched the rates of H4B radical formation and Arg hydroxylation in both enzymes. The rates of H4B radical formation differed such that nNOSoxy (18 s(-1)) > iNOSoxy (11 s(-1)) > eNOSoxy (6 s(-1)), whereas the lifetimes of the resulting H4B radical followed an opposite rank order. 5MeH4B supported a three-fold faster radical formation and greater radical stability relative to H4B in both eNOSoxy and nNOSoxy. Our results indicate the following: (i) the three NOSs share a common mechanism, whereby H4B transfers an electron to the heme-dioxy intermediate. This step enables Arg hydroxylation and is rate-limiting for all subsequent steps in the hydroxylation reaction. (ii) A direct correlation exists between pterin radical stability and the speed of its formation in the three NOSs. (iii) Uncoupled NO synthesis often seen for eNOS at low H4B concentrations may be caused by the slow formation and poor stability of its H4B radical.  相似文献   

17.
It has been established that in the case of inducible NO synthase (NOS), a functionally active homodimer is assembled from the heme-deficient monomeric apo-NOS in vitro by the addition of heme, whereas the heme-deficient neuronal isoform (apo-nNOS) is at best only partially activated. In the current study we have discovered that reactive oxygen species, which can be removed by the addition of superoxide dismutase and catalase, destroy the heme and limit the activation of apo-nNOS in vitro. With the use of these improved conditions, we show for the first time that heme insertion is a rapid process that results in formation of a heme-bound monomeric nNOS that is able to form the ferrous-CO P450 complex but is unable to synthesize NO. A slow process requiring more than 90 min is required for dimerization and activation of this P450 intermediate to give an enzyme with a specific activity of approximately 1100 nmol of NO formed/min/mg of protein, similar to that of the native enzyme. Interestingly, the dimer is not SDS-resistant and is not the same dimer that forms in vivo. These studies indicate at least two intermediates in the assembly of nNOS and advance our understanding of the regulation of nNOS.  相似文献   

18.
Nitric-oxide synthase (NOS) requires the cofactor, (6R)-5,6,7, 8-tetrahydrobiopterin (H4B), for catalytic activity. The crystal structures of NOSs indicate that H4B is surrounded by aromatic residues. We have mutated the conserved aromatic acids, Trp(676), Trp(678), Phe(691), His(692), and Tyr(706), together with the neighboring Arg(414) residue within the H4B binding region of full-length neuronal NOS. The W676L, W678L, and F691L mutants had no NO formation activity and had very low heme reduction rates (<0.02 min(-1)) with NADPH. Thus, it appears that Trp(676), Trp(678), and Phe(691) are important to retain the appropriate active site conformation for H4B/l-Arg binding and/or electron transfer to the heme from NADPH. The mutation of Tyr(706) to Leu and Phe decreased the activity down to 13 and 29%, respectively, of that of the wild type together with a dramatically increased EC(50) value for H4B (30-40-fold of wild type). The Tyr(706) phenol group interacts with the heme propionate and Arg(414) amine via hydrogen bonds. The mutation of Arg(414) to Leu and Glu resulted in the total loss of NO formation activity and of the heme reduction with NADPH. Thus, hydrogen bond networks consisting of the heme carboxylate, Tyr(706), and Arg(414) are crucial in stabilizing the appropriate conformation(s) of the heme active site for H4B/l-Arg binding and/or efficient electron transfer to occur.  相似文献   

19.
20.
M Kohashi  J Tse  W N Piper 《Life sciences》1984,34(2):193-196
Purified rat hepatic uroporphyrinogen (UROgen) I synthase (URO-S) was inhibited by bilirubin or the ditaurine derivative. Inhibition was reversible and non-competitive to the substrate porphobilinogen (PBG). The inhibition constants (Ki values) for bilirubin and the conjugate were 1.5 microM and 0.26 microM respectively. Rats afflicted with hyperbilirubinemia caused by biliary obstruction had decreased levels of hepatic microsomal heme (58% of control) and cytochrome P-450 (60% of control) at day 3. Hepatic delta-aminolevulinic acid synthetase (ALAS) activity was increased (39% of control) at day 3.  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