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 875 毫秒
1.
Jensen FB 《The FEBS journal》2008,275(13):3375-3387
The nitrite reductase activity of deoxyhemoglobin has received much recent interest because the nitric oxide produced in this reaction may participate in blood flow regulation during hypoxia. The present study used spectral deconvolution to characterize the reaction of nitrite with carp and rabbit hemoglobin at different constant oxygen tensions that generate the full range of physiological relevant oxygen saturations. Carp is a hypoxia-tolerant species with very high hemoglobin oxygen affinity, and the high R-state character and low redox potential of the hemoglobin is hypothesized to promote NO generation from nitrite. The reaction of nitrite with deoxyhemoglobin leads to a 1 : 1 formation of nitrosylhemoglobin and methemoglobin in both species. At intermediate oxygen saturations, the reaction with deoxyhemoglobin is clearly favored over that with oxyhemoglobin, and the oxyhemoglobin reaction and its autocatalysis are inhibited by nitrosylhemoglobin from the deoxyhemoglobin reaction. The production of NO and nitrosylhemoglobin is faster and higher in carp hemoglobin with high O(2) affinity than in rabbit hemoglobin with lower O(2) affinity, and it correlates inversely with oxygen saturation. In carp, NO formation remains substantial even at high oxygen saturations. When oxygen affinity is decreased by T-state stabilization of carp hemoglobin with ATP, the reaction rates decrease and NO production is lowered, but the deoxyhemoglobin reaction continues to dominate. The data show that the reaction of nitrite with hemoglobin is dynamically influenced by oxygen affinity and the allosteric equilibrium between the T and R states, and that a high O(2) affinity increases the nitrite reductase capability of hemoglobin.  相似文献   

2.
Nitrite reacts with deoxyhemoglobin to generate nitric oxide (NO). This reaction has been proposed to contribute to nitrite-dependent vasodilation in vivo and potentially regulate physiological hypoxic vasodilation. Paradoxically, while deoxyhemoglobin can generate NO via nitrite reduction, both oxyhemoglobin and deoxyhemoglobin potently scavenge NO. Furthermore, at the very low O(2) tensions required to deoxygenate cell-free hemoglobin solutions in aortic ring bioassays, surprisingly low doses of nitrite can be reduced to NO directly by the blood vessel, independent of the presence of hemoglobin; this makes assessments of the role of hemoglobin in the bioactivation of nitrite difficult to characterize in these systems. Therefore, to study the O(2) dependence and ability of deoxhemoglobin to generate vasodilatory NO from nitrite, we performed full factorial experiments of oxyhemoglobin, deoxyhemoglobin, and nitrite and found a highly significant interaction between hemoglobin deoxygenation and nitrite-dependent vasodilation (P < or = 0.0002). Furthermore, we compared the effect of hemoglobin oxygenation on authentic NO-dependent vasodilation using a NONOate NO donor and found that there was no such interaction, i.e., both oxyhemoglobin and deoxyhemoglobin inhibited NO-mediated vasodilation. Finally, we showed that another NO scavenger, 2-carboxyphenyl-4,4-5,5-tetramethylimidazoline-1-oxyl-3-oxide, inhibits nitrite-dependent vasodilation under normoxia and hypoxia, illustrating the uniqueness of the interaction of nitrite with deoxyhemoglobin. While both oxyhemoglobin and deoxyhemoglobin potently inhibit NO, deoxyhemoglobin exhibits unique functional duality as an NO scavenger and nitrite-dependent NO generator, suggesting a model in which intravascular NO homeostasis is regulated by a balance between NO scavenging and NO generation that is dynamically regulated by hemoglobin's O(2) fractional saturation and allosteric nitrite reductase activity.  相似文献   

3.
Mixtures of nitric oxide and hemoglobin were prepared in a rapid freeze apparatus and analyzed by EPR spectroscopy. Spectra from samples at various degrees of saturation showed that the two subunits bound NO at equal rates. Identical results were observed in 0.1 M phosphate at pH 6.5 and 0.1 M 2,2'-bis(hydroxymethyl)-2,2',2'-nitrilotriethanol, 0.1 M NaCl at pH 7.0, both in the presence and absence of inositol hexaphosphate at either buffer condition. At subsaturating levels of NO (less than 60%), or at all levels of saturation in the presence of inositol hexaphosphate, it was found that the EPR spectrum of nitrosylhemoglobin varied with the length of time before freezing. This change was characterized by the development of a hyperfine structure at g = 2.01 which appeared with a half-time of approximately 0.4 s. Maxwell and Caughey (Maxwell, J. C., and Caughey, W. S. (1976) Biochemistry 15, 388-395) have attributed this three-line EPR hyperfine structure to the formation of a pentacoordinate ferroheme-NO complex. Corresponding slow changes were observed in the visible absorption spectrum following the binding of low levels of NO to deoxyhemoglobin or inositol hexaphosphate to fully saturated nitrosylhemoglobin. Thus it appears that NO binding to the alpha and beta subunits of deoxyhemoglobin takes place at equal rates and, under conditions favoring the T quaternary state (low saturation, presence of inositol hexaphosphate), a further slow structural change takes place, resulting in the cleavage of the iron--proximal histidine bond.  相似文献   

4.
The dissociation of NO from nitrosylhemoglobin   总被引:1,自引:0,他引:1  
The reaction between nitrosylhemoglobin and an excess of deoxymyoglobin has been used to study the kinetics of ligand dissociation from Hb4(NO)4 and Hb4(no)1 species. The kinetics of the dissociation of the first NO molecule from Hb4(no)4 was studied by the ligand replacement method. The results indicate that: (a) the ligand dissociation reaction in Hb4(NO)4 is a cooperative process. This is consistent with the results of Moore and Gibson (Moore, E.G., and Gibson, Q.H. (1977) J. Biol. Chem. 251, 2788-2794). (b) alpha and beta chains in the T state formed by adding IHP to Hb4(NO)4 show kinetic heterogeneity. (c) A similar kinetic heterogeneity is shown by alpha and beta chains in the species Hb4NO in the absence of IHP.(d) The value for the NO dissociation rate constant calculated from the slow phases observed in (b) and (c) is similar to that estimated for the R state. These results suggest that the R to T transition brought about with or without inositol hexaphosphate changes the ligand affinity of one type of the chains much more than of the other. On the basis of IR and EPR studies, it is suggested that alpha chains undergo larger functional changes in R to T transition (or vice versa) in nitrosylhemoglobin. The kinetic parameters for HbNO are compared with those of HbO2 and HbCO and the implications of the results for the reaction mechanism are discussed.  相似文献   

5.
In addition to interacting with hemoglobin as a heme ligand to form nitrosylhemoglobin, NO can react with cysteine sulfhydryl groups to form S-nitrosocysteine or cysteine oxides such as cysteinesulfenic acid. Both modes of interaction are very sensitive to the quaternary structure of hemoglobin. To directly view the interaction of NO with quaternary-T deoxyhemoglobin, crystallographic studies were carried out on crystals of deoxyhemoglobin that were exposed to gaseous NO under a variety of conditions. Consistent with previous spectroscopic studies in solution, these crystallographic studies show that the binding of NO to the heme groups of crystalline wild-type deoxyhemoglobin ruptures the Fe-proximal histidine bonds of the alpha-subunits but not the beta-subunits. This finding supports Perutz's theory that ligand binding induces tension in the alpha Fe-proximal histidine bond. To test Perutz's theory, deoxy crystals of the mutant hemoglobin betaW37E were exposed to NO. This experiment was carried out because previous studies have shown that this mutation greatly reduces the quaternary constraints that oppose the ligand-induced movement of the alpha-heme Fe atom into the plane of the porphyrin ring. As hypothesized, the Fe-proximal histidine bonds in both the beta- and the alpha-subunits remain intact in crystalline betaW37E after exposure to NO. With regard to S-nitrosocysteine or cysteine oxide formation, no evidence for the reaction of NO with any cysteine residues was detected under anaerobic conditions. However, when deoxyhemoglobin crystals are first exposed to air and then to NO, the appearance of additional electron density indicates that Cys93(F9)beta has been modified, most likely to cysteinesulfenic acid. This modification of Cys93(F9)beta disrupts the intrasubunit salt bridge between His146(HC3)beta and Asp94(FG1)beta, a key feature of the quaternary-T hemoglobin structure. Also presented is a reanalysis of our previous crystallographic studies [Chan, N.-L., et al. (1998) Biochemistry 37, 16459-16464] of the interaction of NO with liganded hemoglobin in the quaternary-R2 structure. These studies showed additional electron density at Cys93(F9)beta that was consistent with an NO adduct. However, for reasons discussed in this paper, we now believe that this adduct may be the Hb-S-N.-O-H radical intermediate and not Hb-S-N=O as previously suggested.  相似文献   

6.
NMR was used to study the quaternary structure of nitrosyl- and methemoglobin, the kinetics and equilibrium behavior of nitric oxide binding, and the oxidation of hemoglobin. The -9.6 ppm (from H2O) resonance was used as a measure of nitrosylhemoglobin molecules in the T quaternary structure. We found that stripped nitrosylhemoglobin is 70% in the T state below pH 6.4, and is in the R state above. Inositol hexaphosphate (IHP) raises this transition point to pH 7.5. For stripped aquomethemoglobin, the T marker at -10 ppm is absent. In IHP, at pH 6.5 all of the molecules are in the T state. At both higher and lower pH they shift to the R state. The intensity decreases to half of its maximum at pH 5.5 and 7.4. The relative affinity of nitric oxide binding to the alpha and beta subunits was inferred from the intensities of the resonances at -12 and -18 ppm. Under conditions in which nitrosylhemoglobin exists in the T state, NO binds to the alpha subunit 10 times more strongly than it does to the beta subunit. The kinetic experiments reveal that it binds to the two subunits at the same rate and that it dissociates at 5 x 10(-3) s-1 from the beta subunit and at 5 x 10(-4) s-1 from alpha subunit. At high pH, the two subunits are ligated at the same rate. Potassium ferricyanide oxidation, at pH 6.0 in the absence of IHP, is about 3 times more favorable for the alpha than the beta subunit. Addition of IHP raises this preferential oxidation slightly. At pH 8.44, both alpha and beta subunits were oxidized at the same rate.  相似文献   

7.
Recent evidence suggests that the reaction between nitrite and deoxygenated hemoglobin provides a mechanism by which nitric oxide is synthesized in vivo. This reaction has been previously defined to follow second order kinetics, although variable product stoichiometry has been reported. In this study we have re-examined this reaction and found that under fully deoxygenated conditions the product stoichiometry is 1:1 (methemoglobin:nitrosylhemoglobin), and unexpectedly, the kinetics deviate substantially from a simple second order reaction and exhibit a sigmoidal profile. The kinetics of this reaction are consistent with an increase in reaction rate elicited by heme oxidation and iron-nitrosylation. In addition, conditions that favor the "R" conformation show an increased rate over conditions that favor the "T" conformation. The reactivity of nitrite with heme is clearly more complex than has been previously realized and is dependent upon the conformational state of the hemoglobin tetramer, suggesting that the nitrite reductase activity of hemoglobin is under allosteric control.  相似文献   

8.
Nitrite reacts with deoxyhemoglobin to form nitric oxide (NO) and methemoglobin. Though this reaction is experimentally associated with NO generation and vasodilation, kinetic analysis suggests that NO should not be able to escape inactivation in the erythrocyte. We have discovered that products of the nitrite-hemoglobin reaction generate dinitrogen trioxide (N2O3) via a novel reaction of NO and nitrite-bound methemoglobin. The oxygen-bound form of nitrite-methemoglobin shows a degree of ferrous nitrogen dioxide (Fe(II)-NO2*) character, so it may rapidly react with NO to form N2O3. N2O3 partitions in lipid, homolyzes to NO and readily nitrosates thiols, all of which are common pathways for NO escape from the erythrocyte. These results reveal a fundamental heme globin- and nitrite-catalyzed chemical reaction pathway to N2O3, NO and S-nitrosothiol that could form the basis of in vivo nitrite-dependent signaling. Because the reaction redox-cycles (that is, regenerates ferrous heme) and the nitrite-methemoglobin intermediate is not observable by electron paramagnetic resonance spectroscopy, this reaction has been 'invisible' to experimentalists over the last 100 years.  相似文献   

9.
Recent studies reveal a novel role for hemoglobin as an allosterically regulated nitrite reductase that may mediate nitric oxide (NO)-dependent signaling along the physiological oxygen gradient. Nitrite reacts with deoxyhemoglobin in an allosteric reaction that generates NO and oxidizes deoxyhemoglobin to methemoglobin. NO then reacts at a nearly diffusion-limited rate with deoxyhemoglobin to form iron-nitrosyl-hemoglobin, which to date has been considered a highly stable adduct and, thus, not a source of bioavailable NO. However, under physiological conditions of partial oxygen saturation, nitrite will also react with oxyhemoglobin, and although this complex autocatalytic reaction has been studied for a century, the interaction of the oxy- and deoxy-reactions and the effects on NO disposition have never been explored. We have now characterized the kinetics of hemoglobin oxidation and NO generation at a range of oxygen partial pressures and found that the deoxy-reaction runs in parallel with and partially inhibits the oxy-reaction. In fact, intermediates in the oxy-reaction oxidize the heme iron of iron-nitrosyl-hemoglobin, a product of the deoxy-reaction, which releases NO from the iron-nitrosyl. This oxidative denitrosylation is particularly striking during cycles of hemoglobin deoxygenation and oxygenation in the presence of nitrite. These chemistries may contribute to the oxygen-dependent disposition of nitrite in red cells by limiting oxidative inactivation of nitrite by oxyhemoglobin, promoting nitrite reduction to NO by deoxyhemoglobin, and releasing free NO from iron-nitrosyl-hemoglobin.  相似文献   

10.
The reaction of Fe(II) . bleomycin with O2 to yield Fe(III) . bleomycin has been resolved into two kinetic events by stopped-flow spectrophotometry. The first event is first order with respect to both bleomycin and O2 and may be regarded as a second order reaction (k = 6.1 x 10(3) M-1s-1 at 2 degrees C). The first product has no EPR spectrum. The optical spectrum resembles those of Fe(II) . bleomycin complexes with CO, NO, and ethyl isocyanide. We propose that the first product is an Fe(II) . bleomycin . O2 complex. The second kinetic event is first order with respect to the first accumulated product (k = 0.11 s-1 at 2 degrees C) and independent of oxygen concentration. The product of this reaction is indistinguishable from Fe(III) . bleomycin by optical and EPR spectroscopy.  相似文献   

11.
The formation and disappearance of a photosensitive species during the reaction of reduced cytochrome c oxidase (putatively a3II.O2), EC 1.9.3.1, has been followed by (a) mixing a3II.CO with O2 in a stopped flow apparatus; (b) initiating the oxygen-oxidase reaction by removing CO with a laser flash; (c) probing the reaction mixture for photosensitivity with a second laser flash. Photosensitivity appears in the reaction mixture after the first laser flash, reaches a maximum after 50-60 microseconds ([O2] greater than 100 microM), and disappears in a further 50-100 microseconds. The kinetics can be represented by the scheme [formula: see text]. In species B, O2 is associated with the protein, possibly CuB, but not with the heme. Species C is the photosensitive a3II.O2 complex, and in D, a3 iron has been oxidized. The formation of species C is responsible for the rapid phase of absorbance change in the oxidase-oxygen reaction. The rate of reaction with oxygen approaches the limit of 35,000 s-1 at high oxygen. Nitric oxide, however, reacts with FeII oxidase with a rate of 1 x 10(8) M-1 s-1, which is accurately maintained up to an observed rate of 10(5) s-1. In flash photolysis experiments, approximately half of the photodissociated nitric oxidase recombines in a biphasic geminate reaction with rates of 1 x 10(8) s-1 and 1 x 10(7) s-1.  相似文献   

12.
Widely distributed flavohemoglobins (flavoHbs) function as NO dioxygenases and confer upon cells a resistance to NO toxicity. FlavoHbs from Saccharomyces cerevisiae, Alcaligenes eutrophus, and Escherichia coli share similar spectra, O(2), NO, and CO binding kinetics, and steady-state NO dioxygenation kinetics. Turnover numbers (V(max)) for S. cerevisiae, A. eutrophus, and E. coli flavoHbs are 112, 290, and 365 NO heme(-1) s(-1), respectively, at 37 degrees C with 200 microm O(2). The K(M) values for NO are low and range from 0.1 to 0.25 microm. V(max)/K(M)(NO) ratios of 900-2900 microm(-1) s(-1) indicate an extremely efficient dioxygenation mechanism. Approximate K(M) values for O(2) range from 60 to 90 microm. NO inhibits the dioxygenases at NO:O(2) ratios of > or =1:100 and makes true K(M)(O(2)) values difficult to determine. High and roughly equal second order rate constants for O(2) and NO association with the reduced flavoHbs (17-50 microm(-1) s(-1)) and small NO dissociation rate constants suggest that NO inhibits the dioxygenase reaction by forming inactive flavoHbNO complexes. Carbon monoxide also binds reduced flavoHbs with high affinity and competitively inhibits NO dioxygenases with respect to O(2) (K(I)(CO) = approximately 1 microm). These results suggest that flavoHbs and related hemoglobins evolved as NO detoxifying components of nitrogen metabolism capable of discriminating O(2) from inhibitory NO and CO.  相似文献   

13.
Gas-phase oxidation of nitric oxide: chemical kinetics and rate constant.   总被引:2,自引:0,他引:2  
Inhaled nitric oxide (NO) is gaining popularity as a selective pulmonary vasodilator. Because of the potential toxicity of NO and its oxidizing product nitrogen dioxide (NO2), any system for the delivery of inhaled NO must aim at predictable and reproducible levels of NO and at as low concentrations of NO2 as possible. This review describes the chemical kinetics and rate constant values k for the reaction 2NO + O2 = 2NO2. This reaction has been well established as a third-order homogeneous reaction. Published data support two equally plausible two-step mechanisms for the reaction between NO and O2 over a wide range of temperature and pressure. The Arrhenius equation k (L2 x mol(-2) x s(-1)) = 1.2 x 10(3) e(530/T) (=1.2 x 10(3) x 10(230/T)) gives the best fit to the experimental values of the rate constant thus far reported in a temperature range of 273 to 600 K. Using the reaction mechanism and the rate constant k, one can make reliable predictions about NO2 formation in any set of NO inhalation therapy conditions. It is also pointed out that NO3, the intermediate of one of the two mechanisms, deserves serious attention in NO inhalation therapy.  相似文献   

14.
Production of oxalic acid from sugar beet molasses was developed in a series of three reactors. Nitrogen oxides formed were used to manufacture oxalic acid in the second and third reactor. Parameters affecting the reaction were determined to be, air flow rate, temperature, the amount of V2O5 catalyst and the concentrations of molasses and H2SO4. The maximum yields in the second and third reactors were 78.9% and 74.6% of theoretical yield, respectively. Also, kinetic experiments were performed and the first-order rate constants were determined for the glucose consumption rate. Nitrogen oxides in off-gases from the final reactor were absorbed in water and concentrated sulphuric acid and reused in the following reactors giving slightly lower yields under similar conditions. In this novel way, it was possible to recover NO(x) and to prevent air pollution. Meanwhile, it was possible to reduce the unit cost of reactant for oxalic acid production. A maximum 77.5% and 74.1% of theoretical yield was obtained by using the absorption solutions with NO(x).  相似文献   

15.
The reaction of nitric oxide (*NO) with ubiquinol-0 and ubiquinol-2, short-chain analogs of coenzyme Q, was examined in anaerobic and aerobic conditions in terms of formation of intermediates and stable molecular products. The chemical reactivity of ubiquinol-0 and ubiquinol-2 towards *NO differed only quantitatively, the reactions of ubiquinol-2 being slightly faster than those of ubiquinol-0. The ubiquinol/*NO reaction entailed oxidation of ubiquinol to ubiquinone and reduction of *NO to NO-, the latter identified by its reaction with metmyoglobin to form nitroxylmyoglobin and indirectly by measurement of nitrous oxide (N2O) by gas chromatography. Both the rate of ubiquinone accumulation and *NO consumption were linearly dependent on ubiquinol and *NO concentrations. The stoichiometry of *NO consumed per either ubiquinone formed or ubiquinol oxidized was 1.86 A 0.34. The reaction of *NO with ubiquinols proceeded with intermediate formation of ubisemiquinones that were detected by direct EPR. The second order rate constants of the reactions of ubiquinol-0 and ubiquinol-2 with *NO were 0.49 and 1.6 x 10(4) M(-1)s(-1), respectively. Studies in aerobic conditions revealed that the reaction of *NO with ubiquinols was associated with O2 consumption. The formation of oxyradicals - identified by spin trapping EPR- during ubiquinol autoxidation was inhibited by *NO, thus indicating that the O2 consumption triggered by *NO could not be directly accounted for in terms of oxyradical formation or H2O2 accumulation. It is suggested that oxyradical formation is inhibited by the rapid removal of superoxide anion by *NO to yield peroxynitrite, which subsequently may be involved in the propagation of ubiquinol oxidation. The biological significance of the reaction of ubiquinols with *NO is discussed in terms of the cellular O2 gradients, the steady-state levels of ubiquinols and *NO, and the distribution of ubiquinone (largely in its reduced form) in biological membranes with emphasis on the inner mitochondrial membrane.  相似文献   

16.
Herold S  Röck G 《Biochemistry》2005,44(16):6223-6231
Nitrosylhemoglobin (HbFe(II)NO) has been shown to be generated in vivo from the reaction of deoxyHb with NO(*) as well as with nitrite. Despite the physiological importance attributed to this form of Hb, its reactivity has not been investigated in detail. In this study, we showed that the rate of oxidation of HbFe(II)NO by O(2) does not depend on the O(2) concentration. The reaction time courses had to be fitted to a two-exponential expression, and the obtained rates were approximately 2 x 10(-)(4) and 1 x 10(-)(4) s(-)(1), respectively. In the presence of the allosteric effector inositol hexaphosphate (IHP), the value for the fast component of the rate was significantly larger (44 x 10(-)(4) s(-)(1)) whereas that for the slow step was only slightly higher (2.5 x 10(-)(4) s(-)(1)). Moreover, we found that both in the absence and in the presence of IHP the rate of the O(2)-mediated oxidation of HbFe(II)NO is essentially identical to that of NO(*) dissociation from HbFe(II)NO, determined under analogous conditions by replacement of NO(*) with CO in the presence of an excess of dithionite. Taken together, our data show that the reaction between O(2) and HbFe(II)NO proceeds in three steps via dissociation of NO(*) (rate-determining step), binding of O(2) to deoxyHb, and NO(*)-mediated oxidation of oxyHb to metHb and nitrate.  相似文献   

17.
The widespread opinion that N(2)O(3) as a product of NO oxidation is the only nitros(yl)ating agent under aerobic conditions is based on experiments in homogeneous buffered water solutions. In vivo NO is oxidized in heterogeneous media and this opinion is not correct. The equilibrium in the system being dependent on temperature and DeltaG((sol)) for NO, NO(2), isomers of both N(2)O(3), and N(2)O(4). For polar solvents including water, DeltaG((sol)) for N(2)O(3) is high enough, and a stationary concentration of N(2)O(3) in the mixture with other oxides is sufficient to guarantee the hydrolysis of N(2)O(3) to nitrite. In heterogeneous media, the mixture contains solvates NO(2(sol)), N(2)O(3(sol)), and N(2)O(4(sol)) at stationary nonequilibrium concentrations. As far as DeltaG((sol)) is decreased in heterogeneous mixtures with low polar solvents and/or at increased temperatures, the equilibrium in such a system shifts to NO(2). Although NO(2) is a reactive free radical, it almost does not react with water. In contrast, the reaction with most functional protein groups efficiently proceeds by a radical type with the formation of nitrite and new radicals (X) further stabilized in various forms. Therefore, the ratio of the nitrosylated and nitrated products yields depends on actual concentrations of all NO(x).  相似文献   

18.
The stoichiometry and kinetics of reaction of methemerythrin with the deoxy forms of myoglobin and hemoglobin have been examined at I = 0.2 M and 25 degrees C. One mole of methemerythrin (on the basis of the monomer unit containing two irons) reacts with 2 mol of deoxymyoglobin and with 0.5 mol of deoxyhemoglobin. All reactions are second order. Rate constants for reaction with deoxymyoglobin are 0.25 M-1s-1 (Phascolopsis gouldii) and 5.6 M-1s-1 (Themiste pyroides) at pH 6.3. There is little effect of raising the ionic strength to 1.35 M and only a small decrease in rate when the pH is adjusted to 8.2. The rate constant for reaction of deoxyhemoglobin with P. gouldii methemerythrin is approximately 0.1 M-1s-1 at pH 6.3. Metmyohemerythrin from T. pyroides reacts slightly slower than the octamer form (k = 2.0 M-1s-1 at pH 6.3 and 7.0). Oxymyoglobin is converted to metmyoglobin by methemerythrin. The electron-transfer path is discussed and a self-exchange rate constant for hemerythrin assessed as 10(-3) M-1s-1 on the basis of Marcus's theory.  相似文献   

19.
Hemopexin (HPX), serving as scavenger and transporter of toxic plasma heme, has been postulated to play a key role in the homeostasis of NO. Here, kinetics of HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO are reported. NO reacts reversibly with HPX-heme(II) yielding HPX-heme(II)-NO, according to the minimum reaction scheme: HPX-heme(II)+NO kon<-->koff HPX-heme(II)-NO values of kon, koff, and K (=kon/koff) are (6.3+/-0.3)x10(3)M-1s-1, (9.1+/-0.4)x10(-4)s-1, and (6.9+/-0.6)x10(6)M-1, respectively, at pH 7.0 and 10.0 degrees C. O2 reacts with HPX-heme(II)-NO yielding HPX-heme(III) and NO3-, by means of the ferric heme-bound peroxynitrite intermediate (HPX-heme(III)-N(O)OO), according to the minimum reaction scheme: HPX-heme(II)-NO+O2 hon<--> HPX-heme(III)-N(O)OO l-->HPX-heme(III)+NO3- the backward reaction rate is negligible. Values of hon and l are (2.4+/-0.3)x10(1)M-1s-1 and (1.4+/-0.2)x10(-3)s-1, respectively, at pH 7.0 and 10.0 degrees C. The decay of HPX-heme(III)-N(O)OO (i.e., l) is rate limiting. The HPX-heme(III)-N(O)OO intermediate has been characterized by optical absorption spectroscopy in the Soret region (lambdamax=409 nm and epsilon409=1.51x10(5)M-1cm-1). These results, representing the first kinetic evidence for HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO, might be predictive of transient (pseudo-enzymatic) function(s) of heme carriers.  相似文献   

20.
The reaction of trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) and mitochondria was investigated through differential pulse polarography and fluorimetry. The nitrosyl complex undergoes one-electron reduction centered on the NO ligand site. The reaction between the mitochondrial reductor and trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) exhibits a second order specific rate constant calculated as k=2 x 10(1) M(-1) s(-1). The reduced species, trans-[Ru(NH(3))(4)P(OEt)(3)NO](2+), quickly releases NO, yielding trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+). The low toxicities of both trans-[Ru(NH(3))(4)P(OEt)(3)(NO)](2+) and trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+) and its ability to release NO after reductive activation in a biological medium make the nitrosyl compound a useful model of a hypotensive drug.  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