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 31 毫秒
1.
Nagababu E  Ramasamy S  Rifkind JM 《Biochemistry》2007,46(41):11650-11659
The reaction of nitrite with deoxyhemoglobin (deoxyHb) results in the reduction of nitrite to NO, which binds unreacted deoxyHb forming Fe(II)-nitrosylhemoglobin (Hb(II)NO). The tight binding of NO to deoxyHb is, however, inconsistent with reports implicating this reaction with hypoxic vasodilation. This dilemma is resolved by the demonstration that metastable intermediates are formed in the course of the reaction of nitrite with deoxyHb. The level of intermediates is quantitated by the excess deoxyHb consumed over the concentrations of the final products formed. The dominant intermediate has a spectrum that does not correspond to that of Hb(III)NO formed when NO reacts with methemoglobin (MetHb), but is similar to metHb resulting in the spectroscopic determinations of elevated levels of metHb. It is a delocalized species involving the heme iron, the NO, and perhaps the beta-93 thiol. The putative role for red cell reacted nitrite on vasodilation is associated with reactions involving the intermediate. (1) The intermediate is less stable with a 10-fold excess of nitrite and is not detected with a 100-fold excess of nitrite. This observation is attributed to the reaction of nitrite with the intermediate producing N2O3. (2) The release of NO quantitated by the formation of Hb(II)NO is regulated by changes in the distal heme pocket as shown by the 4.5-fold decrease in the rate constant in the presence of 2,3-diphosphoglycerate. The regulated release of NO or N2O3 as well as the formation of the S-nitroso derivative of hemoglobin, which has also been reported to be formed from the intermediates generated during nitrite reduction, should be associated with any hypoxic vasodilation attributed to the RBC.  相似文献   

2.
Cytochrome cd(1) is a respiratory enzyme that catalyzes the physiological one-electron reduction of nitrite to nitric oxide. The enzyme is a dimer, each monomer containing one c-type cytochrome center and one active site d(1) heme. We present stopped-flow Fourier transform infrared data showing the formation of a stable ferric heme d(1)-NO complex (formally d(1)Fe(II)-NO(+)) as a product of the reaction between fully reduced Paracoccus pantotrophus cytochrome cd(1) and nitrite, in the absence of excess reductant. The Fe-(14)NO nu(NO) stretching mode is observed at 1913 cm(-1) with the corresponding Fe-(15)NO band at 1876 cm(-1). This d(1) heme-NO complex is still readily observed after 15 min. EPR and visible absorption spectroscopic data show that within 4 ms of the initiation of the reaction, nitrite is reduced at the d(1) heme, and a cFe(III) d(1)Fe(II)-NO complex is formed. Over the next 100 ms there is an electron redistribution within the enzyme to give a mixed species, 55% cFe(III) d(1)Fe(II)-NO and 45% cFe(II) d(1)Fe(II)-NO(+). No kinetically competent release of NO could be detected, indicating that at least one additional factor is required for product release by the enzyme. Implications for the mechanism of P. pantotrophus cytochrome cd(1) are discussed.  相似文献   

3.
Recent studies have detected a (1)H nuclear magnetic resonance (NMR) reporter signal of metmyoglobin (metMb) during bradykinin stimulation of an isolated mouse heart. The observation has led to the hypothesis that Mb reacts with cellular nitric oxide (NO). However, the hypothesis depends on an unequivocal detection of metMb signals in vivo. In solution, nitrite oxidization of Mb produces a characteristic set of paramagnetically shifted (1)H NMR signals. In the upfield spectral region, MbO(2) and MbCO exhibit the gammaCH(3) Val E11 signals at -2.8 and -2.4 ppm, respectively. In the same spectral region, nitrite oxidation of Mb produces a set of signals at -3.7 and -4.7 ppm at 35 degrees C. Previous studies have confirmed the visibility of metMb signals in perfused rat myocardium. With bradykinin infusion, perfusion pressure and rate-pressure product decrease, consistent with endogenous NO formation. However, neither myocardial O(2) consumption nor high-energy phosphate levels, as reflected in the (31)P NMR signals, show any significant change. Bradykinin still triggers a similar physiological response even in the presence of CO that is sufficient to inhibit 86% Mb. In all cases, the (1)H NMR spectra from perfused rat myocardium reveal no metMb signals. The results suggest that bradykinin-induced NO does not interact significantly with cellular Mb to produce an NMR-detectable quantity of metMb in the perfused rat myocardium. As a consequence, the experiments cannot confirm the intriguing proposal that Mb acts as a cellular NO scavenger.  相似文献   

4.
Sickle cell disease patients receiving hydroxyurea (HU) therapy have shown increases in the production of nitric oxide (NO) metabolites, which include iron nitrosyl hemoglobin (HbNO), nitrite, and nitrate. However, the exact mechanism by which HU forms HbNO in vivo is not understood. Previous studies indicate that the reaction of oxyhemoglobin (oxyHb) or deoxyhemoglobin (deoxyHb) with HU are too slow to account for in vivo HbNO production. In this study, we show that the reaction of methemoglobin (metHb) with HU to form HbNO could potentially be fast enough to account for in vivo HbNO formation but competing reactions of either excess oxyHb or deoxyHb during the reaction reduces the likelihood that HbNO will be produced from the metHb-HU reaction. Using electron paramagnetic resonance (EPR) spectroscopy we have detected measurable amounts of HbNO and metHb during the reactions of oxyHb, deoxyHb, and metHb with excess hydroxylamine (HA). We also demonstrate HbNO and metHb formation from the reactions of excess oxyHb, deoxyHb, or metHb and HA, conditions that are more likely to mimic those in vivo. These results indicate that the reaction of hydroxylamine with hemoglobin produces HbNO and lend chemical support for a potential role for hydroxylamine in the in vivo metabolism of hydroxyurea.  相似文献   

5.
A resonance Raman band involving significantly the iron(III)-histidine stretching (upsilonFe-His) character is identified for metmyoglobin (metMb) through isotope sensitivity of its low-frequency resonance Raman bands, but the identification was not successful for methemoglobin (metHb) and its isolated alpha and beta subunits. A band at 218 cm-1 of natural abundance metMb exhibited a low-frequency shift for 15N-His-labeled metMb (-1.4 cm-1 shift), while the strong porphyrin bands at 248 and 271 cm-1 did not shift significantly. The frequency of the 218-cm-1 band of metMb decreased by 1.6 cm-1 in D2O, probably due to Ndelta-deuteration of the proximal His, in a similar manner to that of the upsilonFe-His band of deoxyMb in D2O. This 218-cm-1 band shifted slightly to a lower frequency in H2(18)O, whereas it did little upon 54Fe isotopic substitution (<0.3 cm-1), presumably because of the six-coordinate structure. The lack of the 54Fe-isotope shift shows that the 218-cm-1 band is specific to metMb and not due to the deoxy species. The intensity of this band decreased for hydroxymetMb and was indiscernible for cyanometMb. For metHb and its alpha and beta subunits, however, the frequencies of the band around 220 cm-1 were not D2O sensitive. These results suggest an assignment of the band around 220 cm-1 to a pyrrole tilting mode, which significantly contains the Fe-His stretching character for metMb but scarcely for metHb and its subunits. The differences in the isotope sensitivity of this band in different proteins are considered to reflect the heme distortion from the planarity and the Fe-His geometry specific to individual proteins.  相似文献   

6.
Herold S  Shivashankar K 《Biochemistry》2003,42(47):14036-14046
Hemoproteins, in particular, myoglobin and hemoglobin, are among the major targets of peroxynitrite in vivo. The oxygenated forms of these proteins are oxidized by peroxynitrite to their corresponding iron(iii) forms (metMb and metHb). This reaction has previously been shown to proceed via the corresponding oxoiron(iv) forms of the proteins. In this paper, we have conclusively shown that metMb and metHb catalyze the isomerization of peroxynitrite to nitrate. The catalytic rate constants were determined by stopped-flow spectroscopy in the presence and absence of 1.2 mM CO(2) at 20 and 37 degrees C. The values obtained for metMb and metHb, with no added CO(2) at pH 7.0 and 20 degrees C, are (7.7 +/- 0.1) x 10(4) and (3.9 +/- 0.2) x 10(4) M(-1) s(-1), respectively. The pH-dependence of the catalytic rate constants indicates that HOONO is the species that reacts with the iron(iii) center of the proteins. In the presence of 1.2 mM CO(2), metMb and metHb also accelerate the decay of peroxynitrite in a concentration-dependent way. However, experiments carried out at pH 8.3 in the presence of 10 mM CO(2) suggest that ONOOCO(2)(-), the species generated from the reaction of ONOO(-) with CO(2), does not react with the iron(iii) center of Mb and Hb. Finally, we showed that different forms of Mb and Hb protect free tyrosine from peroxynitrite-mediated nitration. The order of efficiency is metMbCN < apoMb < metHb < metMb < ferrylMb < oxyHb < deoxyHb < oxyMb. Taken together, our data show that myoglobin is always a better scavenger than hemoglobin. Moreover, the globin offers very little protection, as the heme-free (apoMb) and heme-blocked (metMbCN) forms only partly prevent nitration of free tyrosine.  相似文献   

7.
The binding of imidazole and imidazole derivatives (4- and 2-MeIm) to the metMb was studied by 1H NMR spectroscopy to elucidate the effects of different methyl substitution positions on the affinity and kinetics of binding to metMb. These exogenous ligands can form stable complexes with metMb except 2-MeIm. Kinetics and equilibrium data for the binding of imidazole and 4-MeIm to metMb have been obtained by 2D EXSY, and the reasons for the different affinity are also discussed.  相似文献   

8.
The heme cofactor in soluble guanylate cyclase (sGC) is a selective receptor for NO, an important signaling molecule in eukaryotes. The sGC heme domain has been localized to the N-terminal 194 amino acids of the beta1 subunit of sGC and is a member of a family of conserved hemoproteins, called the H-NOX family (Heme-Nitric Oxide and/or OXygen-binding domain). Three new members of this family have now been cloned and characterized, two proteins from Legionella pneumophila (L1 H-NOX and L2 H-NOX) and one from Nostoc punctiforme (Np H-NOX). Like sGC, L1 H-NOX forms a 5-coordinate Fe(II)-NO complex. However, both L2 H-NOX and Np H-NOX form temperature-dependent mixtures of 5- and 6-coordinate Fe(II)-NO complexes; at low temperature, they are primarily 6-coordinate, and at high temperature, the equilibrium is shifted toward a 5-coordinate geometry. This equilibrium is fully reversible with temperature in the absence of free NO. This process is analyzed in terms of a thermally labile proximal Fe(II)-His bond and suggests that in both the 5- and 6-coordinate Fe(II)-NO complexes of L2 H-NOX and Np H-NOX, NO is bound in the distal heme pocket of the H-NOX fold. NO dissociation kinetics for L1 H-NOX and L2 H-NOX have been determined and support a model in which NO dissociates from the distal side of the heme in both 5- and 6-coordinate complexes.  相似文献   

9.
Methemoglobin (MetHb), oxyhemoglobin (oxyHb), metmyoglobin (metMb), and oxymyoglobin (oxyMb) catalyze formation of the 7-carboxyheptyl and pentyl radicals from 13-hydroperoxy-9,11-octadecadienoic acid. The relative HPLC-ESR peak height of the pentyl radical to the 7-carboxyheptyl radical was found to depend on the oxygen concentration in the reaction mixture. Under aerobic conditions, the 7-carboxyheptyl radical was predominant for the reaction mixture with ferrous ions (or cytochrome c, metHb, or metMb). On the other hand, under anaerobic conditions, the pentyl radical was predominant for the reaction mixture with ferrous ions (or cytochrome c), but the 7-carboxyheptyl radical was still predominant for the reaction mixture with metHb (or metMb), suggesting that metHb (or metMb) catalyzes the reaction through a mechanism different from that in the case of ferrous ions (or cytochrome c). In order to explain the above results, a mechanism, in which molecular oxygen is not involved, is proposed for the formation of the 7-carboxyheptyl radical in the reaction mixture of 13-HPODE with metHb (or metMb) under anaerobic conditions.  相似文献   

10.
Cytochrome c nitrite reductase catalyzes the six-electron, seven-proton reduction of nitrite to ammonia without release of any detectable reaction intermediate. This implies a unique flexibility of the active site combined with a finely tuned proton and electron delivery system. In the present work, we employed density functional theory to study the recharging of the active site with protons and electrons through the series of reaction intermediates based on nitrogen monoxide [Fe(II)-NO(+), Fe(II)-NO·, Fe(II)-NO(-), and Fe(II)-HNO]. The activation barriers for the various proton and electron transfer steps were estimated in the framework of Marcus theory. Using the barriers obtained, we simulated the kinetics of the reduction process. We found that the complex recharging process can be accomplished in two possible ways: either through two consecutive proton-coupled electron transfers (PCETs) or in the form of three consecutive elementary steps involving reduction, PCET, and protonation. Kinetic simulations revealed the recharging through two PCETs to be a means of overcoming the predicted deep energetic minimum that is calculated to occur at the stage of the Fe(II)-NO· intermediate. The radical transfer role for the active-site Tyr(218), as proposed in the literature, cannot be confirmed on the basis of our calculations. The role of the highly conserved calcium located in the direct proximity of the active site in proton delivery has also been studied. It was found to play an important role in the substrate conversion through the facilitation of the proton transfer steps.  相似文献   

11.
Cytochrome c' (cyt c') from Methylophilus methylotrophus is unusual insofar as the heme has two axial histidine ligands in the oxidized form but one is detached when the protein is reduced. Despite cyt c' having an axial site available for binding small ligands, we show here that only NO binds readily to the ferrous cyt c'. Binding of CO, as well as CN(-), on the other hand requires considerable structural reorganization, or reduction of the disulfide bridge close to the heme. Standard free energies for the binding of NO and CO reveal high selectivity of the ferrous cyt c' for NO, indicating its putative physiological role. In this work, we characterize in detail the kinetics of NO binding and the structural features of the Fe(2+)-NO adduct by stopped-flow and resonance Raman spectroscopy, respectively.  相似文献   

12.
The nitrite anion (NO(-)(2)) has recently received much attention as an endogenous nitric oxide source that has the potential to be supplemented for therapeutic benefit. One major mechanism of nitrite reduction is the direct reaction between this anion and the ferrous heme group of deoxygenated hemoglobin. However, the reaction of nitrite with oxyhemoglobin (oxyHb) is well established and generates nitrate and methemoglobin (metHb). Several mechanisms have been proposed that involve the intermediacy of protein-free radicals, ferryl heme, nitrogen dioxide (NO(2)), and hydrogen peroxide (H(2)O(2)) in an autocatalytic free radical chain reaction, which could potentially limit the usefulness of nitrite therapy. In this study we show that none of the previously published mechanisms is sufficient to fully explain the kinetics of the reaction of nitrite with oxyHb. Based on experimental data and kinetic simulation, we have modified previous models for this reaction mechanism and show that the new model proposed here is consistent with experimental data. The important feature of this model is that, whereas previously both H(2)O(2) and NO(2) were thought to be integral to both the initiation and propagation steps, H(2)O(2) now only plays a role as an initiator species, and NO(2) only plays a role as an autocatalytic propagatory species. The consequences of uncoupling the roles of H(2)O(2) and NO(2) in the reaction mechanism for the in vivo reactivity of nitrite are discussed.  相似文献   

13.
The reaction of nitric oxide (NO) with fast and reduced cytochrome bo(3)(cyt bo(3)) from Escherichia coli has been investigated. The stoichiometry of NO binding to cyt bo(3) was determined using an NO electrode in the [NO] range 1-14 microM. Under reducing conditions, the initial decrease in [NO] following the addition of cyt bo(3) corresponded to binding of 1 NO molecule per cyt bo(3) functional unit. After this "rapid" NO binding phase, there was a slow, but significant rate of NO consumption ( approximately 0.3molNOmol bo(3)(-1)min(-1)), indicating that cyt bo(3) possesses a low level of NO reductase activity. The binding of NO to fast pulsed enzyme was also investigated. The results show that in the [NO] range used (1-14 microM) both fast and pulsed oxidised cyt bo(3) bind NO with a stoichiometry of 1:1 with an observed dissociation constant of K(d)=5.6+/-0.6 microM and that NO binding was inhibited by the presence of Cl(-). The binding of nitrite to the binuclear centre causes spectral changes similar to those observed upon NO binding to fast cyt bo(3). These results are discussed in relation to the model proposed by Wilson and co-workers [FEBS Lett. 414 (1997) 281] where the binding of NO to Cu(B)(II) results in the formation of the nitrosonium (Cu(B)(I)-NO(+)) complex. NO(+) then reacts with OH(-), a Cu(B) ligand, to form nitrite, which can bind at the binuclear centre. This work suggests for the first time that the binding of NO to oxidised cyt bo(3) does result in the reduction of Cu(B).  相似文献   

14.
A previous study of nitrite reduction by Paracoccus pantotrophus cytochrome cd1 at pH 7.0 identified early reaction intermediates. The c-heme rapidly oxidised and nitrite was reduced to NO at the d1-heme. A slower equilibration of electrons followed, forming a stable complex assigned as 55% cFe(III)d1Fe(II)-NO and 45% cFe(II)d1Fe(II)-NO+. No catalytically competent NO release was observed. Here we show that at pH 6.0, a significant proportion of the enzyme undergoes turnover and releases NO. An early intermediate, which was previously overlooked, is also identified; enzyme immediately following product release is a candidate. However, even at pH 6.0 a considerable fraction of the enzyme remains bound to NO so another component is required for full product release. The kinetically stable product formed at the end of the reaction differs significantly at pH 6.0 and 7.0, as does its rate of formation; thus the reaction is critically dependent on pH.  相似文献   

15.
DevS and DosT from Mycobacterium tuberculosis (MTB) are paralogous heme-based sensor kinases that respond to hypoxia and to low concentrations of nitric oxide (NO). Both proteins work with the response regulator DevR as a two-component regulatory system to induce the dormancy regulon in MTB. While DevS and DosT are inactive when dioxygen is bound to the heme Fe(II) at their sensor domain, autokinase activity is observed in their heme Fe(II)-NO counterparts. To date, the conversion between active and inactive states and the reactivity of the heme-oxy complex toward NO have not been investigated. Here, we use stopped-flow UV-vis spectroscopy and rapid freeze quench resonance Raman spectroscopy to probe these reactions in DevS. Our data reveal that the heme-O(2) complex of DevS reacts efficiently with NO to produce nitrate and the oxidized Fe(III) heme through an NO dioxygenation reaction that parallels the catalytic reactions of bacterial flavohemoglobin and truncated hemoglobins. Autophosphorylation activity assays show that the Fe(III) heme state of DevS remains inactive but exhibits a high affinity for NO and forms an Fe(III)-NO complex that is readily reduced by ascorbate, a mild reducing agent. On the basis of these results, we conclude that upon exposure to low NO concentrations, the inactive oxy-heme complex of DevS is rapidly converted to the Fe(II)-NO complex in the reducing environment of living cells and triggers the initiation of dormancy.  相似文献   

16.
Nitrite is present in red blood cells (RBCs) and is proposed to be the largest intravascular storage pool of vasoactive NO. The mechanism by which nitrite exerts NO vasoactivity remains unclear but deoxyHb exhibits nitrite reductase activity. NitrosylHb (HbFe(II)NO) is formed on nitrite reduction by excess deoxyHb, and S-nitrosated Hb (HbSNO) has also been detected in nitrite/deoxyHb incubations. We report data consistent with efficient HbSNO generation from a nitrosylHb intermediate on oxygenation of anaerobic deoxyHb incubations containing physiologically revelant levels of nitrite, whereas previously a labile nitrosylmetHb (HbFe(III)NO) transient was proposed. The HbSNO yield as a function of the initial nitrite concentration varies with the nitrite/deoxyHb ratio, the incubation time, the concentration of added metHb (a nitrite trap), and the concentration of added cyanide (a strong metHb ligand). Our results reveal that metHb strongly attenuates HbSNO formation, which suggests that the met protein may play a regulatory role by limiting the amount of free (or non-Hb-bound) nitrite within RBCs to prevent hypotension.  相似文献   

17.
The reaction of deoxyhemoglobin with nitric oxide (NO) or nitrite ions (NO 2 (-)) produces iron-nitrosyl-hemoglobin (HbNO) in contrast to the reaction with oxyhemoglobin, which produces methemoglobin and nitrate (NO 3 (-)). HbNO has not been associated with the known bioactivities of NO. We hypothesized that HbNO in erythrocytes could be an important source of bioactive NO/nitrite if its oxidation was coupled to the ascorbic acid (ASC) cycle. Studied by absorption and electron paramagnetic resonance (EPR) spectroscopy, DHA oxidized HbNO to methemoglobin and liberated NO from HbNO as determined by chemiluminescence. Both DHA and ascorbate free radical (AFR), the intermediate between ASC and DHA, enhanced NO oxidation to nitrite, but not nitrate; nor did either oxidize nitrite to nitrate. DHA increased the basal levels of nitrite in erythrocytes, while the reactions of nitrite with hemoglobin are slow. In erythrocytes loaded with HbNO, HbNO disappeared after DHA addition, and the AFR signal was detected by EPR. We suggest that the ASC-AFR-DHA cycle may be coupled to that of HbNO-nitrite and provide a mechanism for the endocrine transport of NO via hemoglobin within erythrocytes, resulting in the production of intracellular nitrite. Additionally, intracellular nitrite and nitrate seem to be largely generated by independent pathways within the erythrocyte. These data provide a physiologically robust mechanism for erythrocytic transport of NO bioactivity allowing for hormone-like properties.  相似文献   

18.
Nitric oxide (NO) release from nitric oxide synthases (NOSs) is largely dependent on the dissociation of an enzyme ferric heme-NO product complex (Fe(III)NO). Although the NOS-like protein from Bacillus subtilis (bsNOS) generates Fe(III)NO from the reaction intermediate N-hydroxy-l-arginine (NOHA), its NO dissociation is about 20-fold slower than in mammalian NOSs. Crystal structures suggest that a conserved Val to Ile switch near the heme pocket of bsNOS might determine its kinetic profile. To test this we generated complementary mutations in the mouse inducible NOS oxygenase domain (iNOSoxy, V346I) and in bsNOS (I224V) and characterized the kinetics and extent of their NO synthesis from NOHA and their NO-binding kinetics. The mutations did not greatly alter binding of Arg, (6R)-tetrahydrobiopterin, or alter the electronic properties of the heme or various heme-ligand complexes. Stopped-flow spectroscopy was used to study heme transitions during single turnover NOHA reactions. I224V bsNOS displayed three heme transitions involving four species as typically occurs in wild-type NOS, the beginning ferrous enzyme, a ferrous-dioxy (Fe(II)O(2)) intermediate, Fe(III)NO, and an ending ferric enzyme. The rate of each transition was increased relative to wild-type bsNOS, with Fe(III)NO dissociation being 3.6 times faster. In V346I iNOSoxy we consecutively observed the beginning ferrous, Fe(II)O(2), a mixture of Fe(III)NO and ferric heme species, and ending ferric enzyme. The rate of each transition was decreased relative to wild-type iNOSoxy, with the Fe(III)NO dissociation being 3 times slower. An independent measure of NO binding kinetics confirmed that V346I iNOSoxy has slower NO binding and dissociation than wild-type. Citrulline production by both mutants was only slightly lower than wild-type enzymes, indicating good coupling. Our data suggest that a greater shielding of the heme pocket caused by the Val/Ile switch slows down NO synthesis and NO release in NOS, and thus identifies a structural basis for regulating these kinetic variables.  相似文献   

19.
Recent evidence suggests that the reaction of nitrite with deoxygenated hemoglobin and myoglobin contributes to the generation of nitric oxide and S-nitrosothiols in vivo under conditions of low oxygen availability. We have investigated whether ferrous neuroglobin and cytoglobin, the two hexacoordinate globins from vertebrates expressed in brain and in a variety of tissues, respectively, also react with nitrite under anaerobic conditions. Using absorption spectroscopy, we find that ferrous neuroglobin and nitrite react with a second-order rate constant similar to that of myoglobin, whereas the ferrous heme of cytoglobin does not react with nitrite. Deconvolution of absorbance spectra shows that, in the course of the reaction of neuroglobin with nitrite, ferric Fe(III) heme is generated in excess of nitrosyl Fe(II)-NO heme as due to the low affinity of ferrous neuroglobin for nitric oxide. By using ferrous myoglobin as scavenger for nitric oxide, we find that nitric oxide dissociates from ferrous neuroglobin much faster than previously appreciated, consistently with the decay of the Fe(II)-NO product during the reaction. Both neuroglobin and cytoglobin are S-nitrosated when reacting with nitrite, with neuroglobin showing higher levels of S-nitrosation. The possible biological significance of the reaction between nitrite and neuroglobin in vivo under brain hypoxia is discussed.  相似文献   

20.
Martin E  Berka V  Sharina I  Tsai AL 《Biochemistry》2012,51(13):2737-2746
Soluble guanylyl cyclase (sGC), the key enzyme for the formation of second messenger cyclic GMP, is an authentic sensor for nitric oxide (NO). Binding of NO to sGC leads to strong activation of the enzyme activity. Multiple molecules and steps of binding of NO to sGC have been implicated, but the target of the second NO and the detailed binding mechanism remain controversial. In this study, we used (15)NO and (14)NO and anaerobic sequential mixing-freeze-quench electron paramagnetic resonance to unambiguously confirm that the heme Fe is the target of the second NO. The linear dependence on NO concentration up to 600 s(-1) for the observed rate of the second step of NO binding not only indicates that the binding site of the second NO is different from that in the first step, i.e., the proximal site of the heme, but also supports a concerted mechanism in which the dissociation of the His105 proximal ligand occurs simultaneously with the binding of the second NO molecule. Computer modeling successfully predicts the kinetics of formation of a set of five-coordinate NO complexes with the ligand on either the distal or proximal site and supports the selective release of NO from the distal side of the transient bis-NO-sGC complex. Thus, as has been demonstrated with cytochrome c', a five-coordinate NO-sGC complex containing a proximal NO is formed after the binding of the second 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