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 171 毫秒
1.
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   

2.
NO reactions with hemoglobin (Hb) likely play a role in blood pressure regulation. For example, NO exchange between Hb and S-nitrosoglutathione (GSNO) has been reported in vitro. Here we examine the reaction between GSNO and deoxyHb (HbFe(II)) in the presence of both Cu(I) (2,9-dimethyl-1, 10-phenanthroline (neocuproine)) and Cu(II) (diethylenetriamine-N,N,N',N",N"-pentaacetic acid) chelators using a copper-depleted Hb solution. Spectroscopic analysis of deoxyHb (HbFe(II))/GSNO incubates shows prompt formation (<5 min) of approximately 100% heme-nitrosylated Hb (HbFe(II)NO) in the absence of chelators, 46% in the presence of diethylenetriamine-N,N,N',N",N"-pentaacetic acid, and 25% in the presence of neocuproine. Negligible (<2%) HbFe(II)NO was detected when neocuproine was added to copper-depleted HbFe(II)/GSNO incubates. Thus, HbFe(II)NO formation via a mechanism involving free NO generated by Cu(I) catalysis of GSNO breakdown is proposed. GSH is a source of reducing equivalents because extensive GSSG was detected in HbFe(II)/GSNO incubates in the absence of metal chelators. No S-nitrosation of HbFe(II) was detected under any conditions. In contrast, the NO released from GSNO is directed to Cysbeta(93) of oxyHb in the absence of chelators, but only metHb formation is observed in the presence of chelators. Our findings reveal that the reactions of GSNO and Hb are controlled by copper and that metal chelators do not fully inhibit NO release from GSNO in Hb-containing solutions.  相似文献   

3.
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   

4.
Nitrite reduction to nitric oxide (NO) may be potentiated by a nitrite reductase activity of deoxyHb and contribute to systemic hypoxic vasodilation. The effect of nitrite on the pulmonary circulation has not been well characterized. We explored the effect of nitrite on hypoxic pulmonary vasoconstriction (HPV) and the role of the red blood cell (RBC) in nitrite reduction and nitrite-mediated vasodilation. As to method, isolated rat lungs were perfused with buffer, or buffer with RBCs, and subjected to repeated hypoxic challenges, with or without nitrite. As a result, in buffer-perfused lungs, HPV was reduced at nitrite concentrations of 7 muM and above. Nitrite inhibition of HPV was prevented by excess free Hb and RBCs, suggesting that vasodilation was mediated by free NO. Nitrite-inhibition of HPV was not potentiated by mild acidosis (pH = 7.2) or xanthine oxidase activity. RBCs at 15% but not 1% hematocrit prevented inhibition of HPV by nitrite (maximum nitrite concentration of approximately 35 muM) independent of perfusate Po(2). Degradation of nitrite was accelerated by hypoxia in the presence of RBCs but not during buffer perfusion. In conclusion, low micromolar concentrations of nitrite inhibit HPV in buffer-perfused lungs and when RBC concentration is subphysiological. This effect is lost when RBC concentration approaches physiological levels, despite enhanced nitrite degradation in the presence of RBCs. These data suggest that, although deoxyHb may generate NO from nitrite, insufficient NO escapes the RBC to cause vasodilation in the pulmonary circulation under the dynamic conditions of blood flow through the lungs and that RBCs are net scavengers of NO.  相似文献   

5.
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   

6.
Hemoglobin released in the circulation from ruptured red blood cells can be oxidized by hydrogen peroxide or peroxynitrite to generate the highly oxidizing iron(IV)oxo species HbFe(IV)z=O. Nitrogen monoxide, produced in large amounts by activated inducible nitric oxide synthase, can have indirect cytotoxic effects, mainly through the generation of peroxynitrite from its very fast reaction with superoxide. In the present work we have determined the rate constant for the reaction of HbFe(IV)z=O with NO(*), 2.4 x 10(7) M(-1)s(-1) at pH 7.0 and 20 degrees C. The reaction proceeds via the intermediate HbFe(III)ONO, which then dissociates to metHb and nitrite. As these products are not oxidizing and because of its large rate, the reaction of HbFe(IV)z=O with NO(*) may be important to remove the high valent form of hemoglobin, which has been proposed to be at least in part responsible for oxidative lesions. In addition, we have determined that the rate constant for the reaction of HbFe(IV)z=O with nitrite is significantly lower (7.5 x 10(2) M(-1)s(-1) at pH 7.0 and 20 degrees C), but increases with decreasing pH (1.8 x 10(3) M(-1)s(-1) at pH 6.4 and 20 degrees C). Thus, under acidic conditions as found in ischemic tissues, this reaction may also have a physiological relevance.  相似文献   

7.
Boccini F  Herold S 《Biochemistry》2004,43(51):16393-16404
The strong oxidizing and nitrating agent peroxynitrite has been shown to diffuse into erythrocytes and oxidize oxyhemoglobin (oxyHb) to metHb. Because the value of the second-order rate constant for this reaction is on the order of 10(4) M(-)(1) s(-)(1) and the oxyHb concentration is about 20 mM (expressed per heme), this process is rather fast and oxyHb is considered a sink for peroxynitrite. In this work, we showed that the reaction of oxyHb with peroxynitrite, both in the presence and absence of CO(2), proceeds via the formation of oxoiron(iv)hemoglobin (ferrylHb), which in a second step is reduced to metHb and nitrate by its reaction with NO(2)(*). In the presence of physiological relevant amounts of CO(2), ferrylHb is generated by the reaction of NO(2)(*) with the coordinated superoxide of oxyHb (HbFe(III)O(2)(*)(-)). This reaction proceeds via formation of a peroxynitrato-metHb complex (HbFe(III)OONO(2)), which decomposes to generate the one-electron oxidized form of ferrylHb, the oxoiron(iv) form of hemoglobin with a radical localized on the globin. CO(3)(*)(-), the second radical formed from the reaction of peroxynitrite with CO(2), is also scavenged efficiently by oxyHb, in a reaction that finally leads to metHb production. Taken together, our results indicate that oxyHb not only scavenges peroxynitrite but also the radicals produced by its decomposition.  相似文献   

8.
Nitrosylhemoglobin (HbFe(II)NO) has been detected in vivo, and its role in NO transport and preservation has been discussed. To gain insight into the potential role of HbFe(II)NO, we performed in vitro experiments to determine the effect of oxygenated red blood cells (RBCs) on the dissociation of cell-free HbFe(II)NO, using carboxyhemoglobin (HbFe(II)CO) as a comparison. Results show that the apparent half-life of the cell-free HbFe(II)CO was reduced significantly in the presence of RBCs at 1% hematocrit. In contrast, RBC did not change the apparent half-life of extracellular HbFe(II)NO, but caused a shift in the HbFe(II)NO dissociation product from methemoglobin (metHbFe(III)) to oxyhemoglobin (HbFe(II)O(2)). Extracellular hemoglobin was able to extract CO from HbFe(II)CO-containing RBC, but not NO from HbFe(II)NO-containing RBC. Although these results appear to suggest some unusual interactions between HbFe(II)NO and RBC, the data are explainable by simple HbFe(II)NO dissociation and hemoglobin oxidation with known rate constants. A kinetic model consisting of these reactions shows that (i) deoxyhemoglobin is an intermediate in the reaction of HbFe(II)NO oxidation to metHbFe(III), (ii) the rate-limiting step of HbFe(II)NO decay is the dissociation of NO from HbFe(II)NO, (iii) the magnitude of NO diffusion rate constant into RBC is estimated to be approximately 10(4)M(-1)s(-1), consistent with previous results determined from a competition assay, and (iv) no additional chemical reactions are required to explain these data.  相似文献   

9.
Elucidating the reaction of nitric oxide (NO) with oxyhemoglobin [HbFe(II)O2] is critical to understanding the metabolic fate of NO in the vasculature. At low concentrations of NO, methemoglobin [HbFe(III)] is the only detectable product from this reaction; however, locally high concentrations of NO have been demonstrated to result in some iron-nitrosylhemoglobin [HbFe(II)NO] and S-nitrosohemoglobin (SNO-Hb) formation. Reductive nitrosylation through a HbFe(III) intermediate was proposed as a viable pathway under such conditions. Here, we explore another potential mechanism based on mixed valenced Hb tetramers. The oxidation of one or two heme Fe(II) in the R-state HbFe(II)O2 has been observed to lower the oxygen affinity of the remaining heme groups, thus creating the possibility of oxygen release and NO binding at the heme Fe(II) sites. This mixed valenced hypothesis requires an allosteric transition of the Hb tetramer. Hence, this hypothesis can account for HbFe(II)NO formation, but not SNO-Hb formation. Here, we demonstrate that cyanide attenuated the formation of SNO-Hb by 30-40% when a saturated NO bolus was added to 0.1-1.0 mM HbFe(II)O2 solutions. In addition, HbFe(II)NO formation under such inhomogeneous conditions does not require allostericity. Therefore, we concluded that the mixed valenced theory does not play a major role under these conditions, and reductive nitrosylation accounts for a significant fraction of the HbFe(II)NO formed and approximately 30-40% of SNO-Hb. The remaining SNO-Hb is likely formed from NO oxidation products.  相似文献   

10.
Nitric oxide (NO) plays a crucial role in human physiology by regulating vascular tone and blood flow. The short life-span of NO in blood requires a mechanism to retain NO bioactivity in the circulation. Recent studies have suggested a mechanism involving the reduction of nitrite back to NO by deoxyhemoglobin in RBCs. A role for RBCs in transporting NO must, however, bypass the scavenging of NO in RBCs by hemoglobin. To understand how the nitrite reaction can deliver bioactive NO to the vasculature, we have studied the intermediates formed during the reaction. A reliable measure of the total concentration of heme-associated nitrite/NO intermediates formed was provided by combining filtration to measure free nitrite by chemiluminescence and electron paramagnetic resonance to measure the final product Hb(II)NO. By modifying the chemiluminescence method used to detect NO, we have been able to identify two intermediates: 1) a heme-associated nitrite complex that is released as NO in acid solution in the presence of ascorbate and 2) an intermediate that releases NO at neutral pH in the presence of ferricyanide when reacted with an Fe(III) ligand like azide. This species designated as “Hb(II)NO+ ⇆ Hb(III)NO” has properties of both isomeric forms resulting in a slower NO dissociation rate and much higher stability than Hb(III)NO, but provides a potential source for bioactive NO, which can be released from the RBC. This detailed analysis of the nitrite reaction with deoxyHb provides important insights into the mechanism for nitrite induced vasodilation by RBCs.Nitric oxide (NO), also known as the endothelium-derived relaxing factor, is an important messenger molecule involved in the regulation of vascular tone and blood flow (1). The primary source for the synthesis of NO in the circulatory system involves endothelial nitric-oxide synthase (2). This enzyme requires oxygen for the synthesis of NO and is, therefore, less effective in the microcirculation where hypoxic vasodilation regulates the delivery of oxygen. Because nitric oxide has a life-time in blood of <2 ms (3), a mechanism is required to allow for more distal and sustained effects of NO at the reduced oxygen pressures found in the microcirculation. Recent studies have suggested that the bioactivity of NO can be conserved in the blood by the uptake of NO and/or nitrite by red blood cells (RBCs)2 and its interaction with hemoglobin (47). However, any role for the red cell in transporting nitric oxide must be able to avoid the very efficient scavenging of nitric oxide by both oxyhemoglobin (oxyHb) and deoxyhemoglobin (deoxyHb) that destroy and trap NO, respectively, preventing a physiological role for RBC NO.In a series of studies, Stamler and co-workers (710) have hypothesized that NO can bypass this difficulty by being transferred to the β-93 thiol group of hemoglobin (Hb) forming S-nitrosylated hemoglobin (SNO-Hb) when partially heme nitrosylated hemoglobin (Hb(II)NO) is oxygenated. The allosteric quaternary conformational change of hemoglobin at low oxygen pressure destabilizes the β-93 nitrosylated thiol and results in the transfer of NO to membrane thiol groups facilitating the release of the NO to the plasma and the vasculature. However, the extremely low levels of SNO-Hb (11) found in human blood and its instability (12) as a result of intracellular reducing conditions within the RBCs do not support the SNO-Hb hypothesis as the major mechanism for NO transport (1113).The 2003 studies by Rifkind and Gladwin and their collaborators (4, 5, 14, 15) proposed an alternative mechanism that involved the reduction of nitrite, formed by the oxidation of NO, back to NO by a reaction with deoxyHb. Nitrite is present in the blood at fairly high levels (0.1–0.5 μmol/liter) (4, 1618), and it is much more stable than NO or S-nitrosothiols (6), making nitrite an ideal storage pool that can be converted to NO. However, the mechanism by which the NO produced in the red cell by nitrite reduction is exported without being trapped or destroyed is still unclear. Recent studies by Rifkind and co-workers (5, 13, 19) have suggested that the trapping of NO by deoxyHb and/or oxyHb can be bypassed by the formation of a metastable intermediate(s) that retains the NO in a state that is not quenched by reacting with oxyHb or deoxyHb.In this report, we quantitate the two intermediate species that are formed during the reduction of nitrite by deoxyHb when an excess of hemoglobin is present. We also demonstrate that one of the intermediate species designated as “Hb(II)NO+ ⇆ Hb(III)NO” has properties of Hb(II)NO+ and Hb(III)NO, respectively. This species has a slower NO dissociation rate and a much higher stability than Hb(III)NO. This intermediate is a potential source for bioactive NO that can be released from RBCs.  相似文献   

11.
Nitrite binds reversibly to the ferriheme proteins metmyoglobin and methemoglobin in aqueous buffer solution at a physiological pH of 7.4. The spectral changes recorded for the formation of metMb(NO2-) differ significantly from those observed for the nitrosylation of metMb, which can be accounted for in terms of the different reaction products. Nitric oxide binding to metMb produces a nitrosyl product with Fe(II)-NO+ character, whereas the reaction with nitrite produces an Fe(III)-NO2- complex. The kinetics of the binding and release of nitrite by metMb and metHb were investigated by stopped-flow techniques at ambient and high pressure. The kinetic traces recorded for the reaction of nitrite with metMb exhibit excellent single-exponential fits, whereas nitrite binding to metHb is characterized by double-exponential kinetics which were assigned to the reactions of the alpha- and beta-chains of metHb with NO2-. The rate constants for the binding of nitrite to metMb and metHb were found to be much smaller than those reported for the binding of NO, such that nitrite impurities will not affect the latter reaction. The activation parameters (deltaH++,deltaS(ne),deltaV++) obtained from the temperature and pressure dependence of the reactions support the operation of a dissociative mechanism for the binding and release of nitrite, similar to that found for the binding and release of NO in metMb.  相似文献   

12.
The discovery that hemoglobin (Hb) in erythrocytes contains a fraction of beta-Cys-93 thiols as the nitrosylated derivative (HbSNO) led to the suggestion that this species is involved in transporting and releasing nitric oxide, which is the signal for local vasodilation. The release of NO from HbSNO requires an electron transfer to facilitate release and to regenerate the cysteine thiol via one-electron reduction in the absence of added thiols. An alternative mechanism, which has received much attention, transfers the nitrosyl group to an external thiol, which in turn would have to be reduced. The observed first order rate constant for the spontaneous oxidation of the ferrous heme of deoxy HbSNO is 1.0 x 10(-4)s(-1) in the absence of thiols. Under the same conditions, native Hb is stable. The oxidation of HbSNO occurs with the same rate constant that can be derived for the rate reported for the formation of HbNO from HbSNO. These similarities suggest that both processes involve the same reaction: internal electron transfer and direct release of nitric oxide.  相似文献   

13.
We previously demonstrated that inhaling nitric oxide (NO) increases the oxygen affinity of sickle red blood cells (RBCs) in patients with sickle cell disease (SCD). Our recent studies found that NO lowered the P50 values of sickle hemoglobin (HbS) hemolysates but did not increase methemoglobin (metHb) levels, supporting the role of NO, but not metHb, in the oxygen affinity of HbS. Here we examine the mechanism by which NO increases HbS oxygen affinity. Because anti-sickling agents increase sickle RBC oxygen affinity, we first determined whether NO exhibits anti-sickling properties. The viscosity of HbS hemolysates, measured by falling ball assays, increased upon deoxygenation; NO treatment reduced the increment. Multiphoton microscopic analyses showed smaller HbS polymers in deoxygenated sickle RBCs and HbS hemolysates exposed to NO. These results suggest that NO inhibits HbS polymer formation and has anti-sickling properties. Furthermore, we found that HbS treated with NO exhibits an isoelectric point similar to that of HbA, suggesting that NO alters the electric charge of HbS. NO–HbS adducts had the same elution time as HbA upon high performance liquid chromatography analysis. This study demonstrates that NO may disrupt HbS polymers by abolishing the excess positive charge of HbS, resulting in increased oxygen affinity.  相似文献   

14.
Transfusion of stored red blood cells (RBCs) is associated with increased morbidity and mortality in trauma patients. Pro-oxidant, pro-inflammatory, and nitric oxide (NO) scavenging properties of stored RBCs are thought to underlie this association. In this study we determined the effects of RBC washing and nitrite and antiheme therapy on stored RBC-dependent toxicity in the setting of trauma-induced hemorrhage. A murine (C57BL/6) model of trauma–hemorrhage and resuscitation with 1 or 3 units of RBCs stored for 0–10 days was used. Tested variables included washing RBCs to remove lower MW components that scavenge NO, NO-repletion therapy using nitrite, or mitigation of free heme toxicity by heme scavenging or preventing TLR4 activation. Stored RBC toxicity was determined by assessment of acute lung injury indices (airway edema and inflammation) and survival. Transfusion with 5 day RBCs increased acute lung injury indexed by BAL protein and neutrophil accumulation. Washing 5 day RBCs prior to transfusion did not decrease this injury, whereas nitrite therapy did. Transfusion with 10 day RBCs elicited a more severe injury resulting in ~90% lethality, compared to <15% with 5 day RBCs. Both washing and nitrite therapy significantly protected against 10 day RBC-induced lethality, suggesting that washing may be protective when the injury stimulus is more severe. Finally, a spectral deconvolution assay was developed to simultaneously measure free heme and hemoglobin in stored RBC supernatants, which demonstrated significant increases of both in stored human and mouse RBCs. Transfusion with free heme partially recapitulated the toxicity mediated by stored RBCs. Furthermore, inhibition of TLR4 signaling, which is stimulated by heme, using TAK-242, or hemopexin-dependent sequestration of free heme significantly protected against both 5 day and 10 day mouse RBC-dependent toxicity. These data suggest that RBC washing, nitrite therapy, and/or antiheme and TLR4 strategies may prevent stored RBC toxicities.  相似文献   

15.
Storage of erythrocytes in blood banks is associated with biochemical and morphological changes to RBCs (red blood cells). It has been suggested that these changes have potential negative clinical effects characterized by inflammation and microcirculatory dysfunction which add to other transfusion-related toxicities. However, the mechanisms linking RBC storage and toxicity remain unclear. In the present study we tested the hypothesis that storage of leucodepleted RBCs results in cells that inhibit NO (nitric oxide) signalling more so than younger cells. Using competition kinetic analyses and protocols that minimized contributions from haemolysis or microparticles, our data indicate that the consumption rates of NO increased ~40-fold and NO-dependent vasodilation was inhibited 2-4-fold comparing 42-day-old with 0-day-old RBCs. These results are probably due to the formation of smaller RBCs with increased surface area: volume as a consequence of membrane loss during storage. The potential for older RBCs to affect NO formation via deoxygenated RBC-mediated nitrite reduction was also tested. RBC storage did not affect deoxygenated RBC-dependent stimulation of nitrite-induced vasodilation. However, stored RBCs did increase the rates of nitrite oxidation to nitrate in vitro. Significant loss of whole-blood nitrite was also observed in stable trauma patients after transfusion with 1 RBC unit, with the decrease in nitrite occurring after transfusion with RBCs stored for >25?days, but not with younger RBCs. Collectively, these data suggest that increased rates of reactions between intact RBCs and NO and nitrite may contribute to mechanisms that lead to storage-lesion-related transfusion risk.  相似文献   

16.
A saturated nitric oxide (NO) solution (1.88 mM) infused i.v. in the anesthetized pig at a dose of 68 nmol/kg/min for 24 min resulted in a time-dependent increase of nitrosylhemoglobin [HbFe(II)NO] as determined by electron spin resonance (ESR), reaching a C(max) of 7.99 +/- 0.42 microM at the end of the infusion, compared to 1.13 +/- 0.42 microM before (p < 0.01). This indicates that NO i.v. is efficiently bioconserved as HbFe(II)NO (approximately 34% of the NO dose) and to a greater extent than by the oxidative pathway (approximately 24% of the NO dose), as determined by measuring plasma nitrites/nitrates (chemiluminescence) and Met-Hb (ESR analysis). When the NO infusion was stopped, HbFe(II)NO declined with a t(1/2) of 15 min, indicating that it is a stable storage form of NO, able to deliver NO distally to the site of administration. No significant differences were observed in systemic and pulmonary vascular resistances during and after NO infusion, but PO(2) showed a significant decrease 15 and 30 min after the infusion. Thus, in normoxic/physiological conditions, HbFe(II)NO does not induce significant NO-dependent vasorelaxation.  相似文献   

17.
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   

18.
Bound NO in human red blood cells: fact or artifact?   总被引:1,自引:0,他引:1  
There has been considerable debate over the nature and chemistry of the interaction between nitric oxide (NO) and red blood cells (RBCs), in particular whether hemoglobin consumes or conserves NO bioactivity. Given the vast range of nitrosation levels reported for human RBCs in the literature, we sought to investigate whether there was a common denominator that could account for such discrepancies across different methodologies and reaction conditions and if such a pathway may exist in physiology. Here, we show that there are marked differences in reactivity toward NO between human and rat hemoglobin, which offers a mechanistic explanation for why basal levels of NO-adducts in primate RBCs are considerably lower than those in rodents. We further demonstrate that the inadvertent introduction of trace amounts of nitrite and incomplete thiol alkylation lead to rapid heme and thiol nitros(yl)ation, with generation of nitrosylhemoglobin (NOHb) and S-nitrosohemoglobin (SNOHb), while neither species is detectable in human RBCs at physiological nitrite concentrations. Thus, caution should be exercised in interpreting experimental results on SNOHb/NOHb levels that were obtained in the absence of knowledge about the degree of nitrite contamination, in particular when a physiological role for such species is implicated.  相似文献   

19.
In this paper, inactivation of alcohol dehydrogenase (ADH) by products of reactions of H2O2 with metHb has been studied. Inactivation of the enzyme was studied in two systems corresponding to two kinetic stages of the reaction. In the first system H2O2 was added to the mixture of metHb and ADH [the (metHb+ADH)+H2O2] system (ADH was present in the system since the moment of addition of H2O2 i. e. since the very beginning of the reaction of metHb with H2O2). In the second system ADH was added to the system 5 min after the initiation of the reaction of H2O2 with metHb [the (metHb+H2O2)5 min+ADH] system. In the first case all the products of reaction of H2O2 with metHb (non-peroxyl and peroxyl radicals and non-radical products, viz. hydroperoxides and *HbFe(IV)=O) could react with the enzyme causing its inactivation. In the second system, enzyme reacted almost exclusively with non-radical products (though a small contribution of reactions with peroxyl radicals cannot be excluded). ADH inactivation was observed in both system. Hydrogen peroxide alone did not inactivate ADH at the concentrations employed evidencing that enzyme inactivation was due exclusively to products of reaction of H2O2 with metHb. The rate and extent of ADH inactivation were much higher in the first than in the second system. The dependence of ADH activity on the time of incubation with ferryl derivatives of Hb can be described by a sum of three exponentials in the first system and two exponentials in the second system. Reactions of appropriate forms of the ferryl derivatives of hemoglobin have been tentatively ascribed to these exponentials. The extent of the enzyme inactivation in the second system was dependent on the proton concentration, being at the highest at pH 7.4 and negligible at pH 6.0. The reaction of H2O2 with metHb resulted in the formation of cross-links of Hb subunits (dimers and trimers). The amount of the dimers formed was much lower in the first system i. e. when the radical forms dominated the reaction of inactivation.  相似文献   

20.
The reaction between hemoglobin (Hb) and NO* has been investigated thoroughly in recent years, but its mechanism is still a matter of substantial controversy. We have carried out a systematic study of the influence of the following factors on the yield of S-nitrosohemoglobin (SNO-Hb) generated from the reaction of NO* with oxy-, deoxy-, and metHb: 1) the volumetric ratio of the protein and the NO* solutions; 2) the rate of addition of the NO* solution to the protein solution; 3) the amount of NO* added; and 4) the concentration of the phosphate buffer. Our results suggest that the highest SNO-Hb yields are mostly obtained by very slow addition of substoichiometric amounts of NO* from a diluted solution. Possible pathways of SNO-Hb formation from the reaction of NO* with oxy-, deoxy-, and metHb are described. Our data strongly suggest that, because of mixing artifacts, care should be taken to use results from in vitro experiments to draw conclusion on the mechanism of the reaction in vivo.  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