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 703 毫秒
1.
Hemoglobin (Hb) vesicles have been developed as cellular-type Hb-based O(2) carriers in which a purified and concentrated Hb solution is encapsulated with a phospholipid bilayer membrane. Ferrous Hb molecules within an Hb vesicle were converted to ferric metHb by reacting with reactive oxygen species such as hydrogen peroxide (H(2)O(2)) generated in the living body or during the autoxidation of oxyHb in the Hb vesicle, and this leads to the loss of O(2) binding ability. The prevention of metHb formation by H(2)O(2) in the Hb vesicle is required to prolong the in vivo O(2) carrying ability. We found that a mixed solution of metHb and L-tyrosine (L-Tyr) showed an effective H(2)O(2) elimination ability by utilizing the reverse peroxidase activity of metHb with L-Tyr as an electron donor. The time taken for the conversion of half of oxyHb to metHb (T(50)) was 420 min for the Hb vesicles containing 4 g/dL (620 microM) metHb and 8.5 mM L-Tyr ((metHb/L-Tyr) Hb vesicles), whereas the time of conversion for the conventional Hb vesicles was 25 min by stepwise injection of H(2)O(2) (310 microM) in 10 min intervals. Furthermore, in the (metHb/L-Tyr) Hb vesicles, the metHb percentage did not reach 50% even after 48 h under a pO(2) of 40 Torr at 37 degrees C, whereas T(50) of the conventional Hb vesicles was 13 h under the same conditions. Moreover, the T(50) values of the conventional Hb vesicles and the (metHb/L-Tyr) Hb vesicles were 14 and 44 h, respectively, after injection into rats (20 mL/kg), confirming the remarkable inhibitory effect of metHb formation in vivo in the (metHb/L-Tyr) Hb vesicles.  相似文献   

2.
Methemoglobin (metHb) with H2O2 catalyzed the oxidation of 3-hydroxykynurenine (3-HKY) in the reaction mixture of metHb, 3-HKY, and H2O2. The spectrophotometric experiments suggest the following mechanism for the 3-HKY oxidation by metHb with H2O2. MetHb first reacts with H2O2 to form the ferryl complex of Hb. This species then oxidizes 3-HKY, while it returns to metHb. 3-HKY was more reactive with the ferryl complex than glutathione but less reactive than ascorbic acid. Scavengers of the hydroxyl radical, dimethyl sulfoxide and ethanol, scarcely inhibited the 3-HKY oxidation by metHb with H2O2. Desferrioxamine, a metal chelator, hardly suppressed the 3-HKY oxidation. These results indicate that the hydroxyl radical is not involved in the 3-HKY oxidation by metHb with H2O2.  相似文献   

3.
Double mixing stopped-flow experiments have been performed to study the stability of asymmetric hemoglobin (Hb) hybrids, consisting of a deoxy and a liganded dimer. The doubly liganded [deoxy/cyano-met] hybrid (species 21) was reported to have an enhanced stability, with tetramer to dimer dissociation requiring over 100 seconds, based on a method that required an incubation of over two days. However, kinetic experiments revealed rapid ligand binding to species 21, as for triply liganded tetramers, which dissociate within a few seconds. For the present study, [deoxy dimer/azido-met dimer] hybrids are formed within 200 ms by stopped-flow mixing of dithionite with a solution containing oxyHb and azido-metHb. The dithionite scavenges oxygen, thus transforming oxyHb to deoxyHb, and the [oxy dimer/azido-met dimer] hybrid to the asymmetric [deoxy/azido-met] hybrid (species 21). After a variable aging time of the asymmetric hybrids, their allosteric state is probed by CO binding in a second mixing. As previously observed the freshly produced asymmetric hybrids bind CO rapidly as for R-state Hb. As the hybrids are aged from 0.1 to 10 seconds, the fraction of slow CO binding increases, consistent with a dissociation of the asymmetric hybrid to form the more stable deoxy Hb tetramer which reacts slowly with CO. Control experiments showed a predominantly slow phase for deoxy Hb, and fast rebinding for the symmetric hybrids.The kinetic data can be simulated with a tetramer to dimer dissociation rate for species 21 of 1.5/second at 100 mM NaCl (pH 7.2) and 1.9/second at 180 mM NaCl (pH 7.4). These values are similar to those reported for liganded Hb, as opposed to deoxy (T-state) tetramers which dissociate over four orders of magnitude more slowly. As expected from simulations of dimer exchange, the observed transition rate depends on the initial fractions of oxy- and metHb; this effect is not consistent with a slow R to T transition. These results, showing a lifetime of about one second for species 21, do not support the symmetry rule which is based on an enhanced stability of the asymmetric hybrid.  相似文献   

4.
Hemoglobin (Hb) has been widely studied as a raw material for various types of oxygen carriers. In the purification of Hb from red blood cells including virus inactivation and denaturation of other proteins and the long-term storage of Hb vesicles (HbV), a deoxygenation process is one of the important processes because of the high stability of deoxygenated Hb to heating and metHb formation. Though an oxygenated Hb solution can be deoxygenated with an artificial lung, it is difficult to reduce the oxygen partial pressure of the Hb solution to less than 10 Torr. We developed an electrochemical system for complete deoxygenation of the Hb solution at the cathode compartment using hydrogen containing nitrogen gas at the anode compartment. Oxygen in the Hb solution was reduced to OH(-) at the cathode compartment within several minutes at a potential value of -1.67 V and was finally converted to water by neutralization with H(+) from the anode in the whole system. The resulting completely deoxygenated Hb could tolerate heat treatment at 62 degrees C for 10 h with no denaturation of deoxygenated Hb. The metHb formation rate of reoxygenated Hb at 37 degrees C was not changed after heat treatment. Furthermore, vesicular stomatitis virus (VSV) could be inactivated at an inactivation degree of more than 5.96 log by heat treatment.  相似文献   

5.
Previous studies showed that CO/H2O oxidation provides electrons to drive the reduction of oxidized hemoglobin (metHb). We report here that Cu(II) addition accelerates the rate of metHb beta chain reduction by CO by a factor of about 1000. A mechanism whereby electron transfer occurs via an internal pathway coupling CO/H2O oxidation to Fe(III) and Cu(II) reduction is suggested by the observation that the copper-induced rate enhancement is inhibited by blocking Cys-beta93 with N-ethylmaleimide. Furthermore, this internal electron-transfer pathway is more readily established at low Cu(II) concentrations in Hb Deer Lodge (beta2His --> Arg) and other species lacking His-beta2 than in Hb A0. This difference is consistent with preferential binding of Cu(II) in Hb A0 to a high affinity site involving His-beta2, which is ineffective in promoting electron exchange between Cu(II) and the beta heme iron. Effective electron transfer is thus affected by Hb type but is not governed by the R left arrow over right arrow T conformational equilibrium. The beta hemes in Cu(II)-metHb are reduced under CO at rates close to those observed for cytochrome c oxidase, where heme and copper are present together in the oxygen-binding site and where internal electron transfer also occurs.  相似文献   

6.
Hemoglobin (Hb) vesicles (particle diameter, ca. 250 nm) have been developed as Hb-based oxygen carriers in which a purified Hb solution is encapsulated with a phospholipid bilayer membrane. The oxidation of Hb to nonfunctional ferric Hb (metHb) was caused by reactive oxygen species, especially hydrogen peroxide (H(2)O(2)), in vivo in addition to autoxidation. We focused on the enzymatic elimination of H(2)O(2) to suppress the metHb formation in the Hb vesicles. In this study, we coencapsulated catalase with Hb within vesicles and studied the rate of metHb formation in vivo. The Hb vesicles containing 5.6 x 10(4) unit mL(-1) catalase decreased the rate of metHb formation by half in comparison with Hb vesicles without catalase. We succeeded in prolonging the oxygen-carrying ability of the Hb vesicle in vivo by the coencapsulation of catalase.  相似文献   

7.
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   

8.
We report the use of electrospray ionization (ESI) mass spectrometry (MS) in conjunction with online rapid mixing to monitor the kinetics of acid-induced ferrihemoglobin denaturation. Under equilibrium conditions, the hemoglobin mass spectrum is dominated by the intact heterotetramer. Dimeric and monomeric species are also observed at lower intensities. In addition, ionic signals corresponding to hexameric (tetramer-dimer) and octameric (tetramer x 2) hemoglobin species are observed. These complexes may represent weak solution-phase assemblies. The acid-induced denaturation process was monitored for reaction time ranging from 9 ms to approximately 3 s. The data obtained were subjected to a global analysis procedure which simultaneously fit all kinetic (ESI-MS intensity vs time) profiles to multiexponential expressions. Results of the global analysis are consistent with the coexistence of two subpopulations of tetrameric hemoglobin which differ in their disassembly rates and ESI charge states. The higher-charge state tetramer ions preferentially dissociate via a rapid pathway (tau(1) = 51 ms), resulting in the transient formation of a heme-saturated dimer, holo-alpha-globin, and a heme-deficient dimer. The latter is shown by MS/MS to be comprised of a heme-bound alpha-subunit complexed with an apo-beta-chain. The slow-decaying tetramer population, apparent at a slightly lower average charge state, breaks down into its monomeric constituents with no observable intermediate species (tau(2) = 390 ms). Surprisingly, unfolded apo-alpha-globin is formed more rapidly than unfolded apo-beta-globin. The appearance of the latter occurs with a relaxation time tau(3) of 1.2 s. It is postulated that accumulation of unfolded apo-beta-globin is delayed by transient population of an undetected unfolding intermediate.  相似文献   

9.
Posttranslational modifications (PTMs) have been reported in hemoglobin (Hb) treated with ROS/RNS in cell-free experiments. However, little is known about oxidative PTMs of Hb occurring within the erythrocytes. The aim of this study is to characterize the patterns of Hb PTMs in erythrocytes under oxidative stress. Using mass spectrometry, we investigated specifically methionine/tryptophan oxidation, tyrosine nitration, and the modification via 4-hydroxynonenal (HNE), a product of lipid-peroxidation, on Hb. We demonstrated that the treatment with H2O2/nitrite induced higher levels of Hb oxidation/nitration in purified Hb preparations than in unpurified hemolysates and erythrocytes, indicating that ROS/RNS are primarily removed by antioxidative mechanisms. We further studied Hb from erythrocytes exposed to γ-irradiation. An irradiation of 30–100 Gy triggered a remarkable increase of intracellular ROS. However, 30 Gy did not induce apparent changes in Hb oxidation/nitration and hemolysis, while Hb oxidation/nitration and hemolysis were significantly enhanced by 100 Gy, suggesting that Hb oxidation/nitration are the consequence of overwhelmed antioxidative mechanisms after oxidative attack and reflect the severity of the oxidative damage of erythrocytes. Although irradiation was known to induce lipid-peroxidation, we could not detect HNE-Hb adducts in irradiated erythrocytes. Analyzing PTM patterns suggests Hb nitration as a more suitable indicator of the oxidative damage of erythrocytes.  相似文献   

10.
Ferric methemoglobin is reduced to its ferrous form by photoirradiation either by direct photoexcitation of the heme portion to induce electron transfer from the surrounding media (Sakai at al. (2000) Biochemistry 39, 14595-14602) or by an indirect electron transfer from a photochemically reduced electron mediator such as flavin. In this research, we studied the mechanism and optimal condition that facilitates photoreduction of flavin mononucleotide (FMN) to FMNH(2) by irradiation of visible light, and the succeeding reduction of concentrated metHb in phospholipid vesicles to restore its O(2) binding ability. Visible light irradiation (435 nm) of a metHb solution containing FMN and an electron donor such as EDTA showed a significantly fast reduction to ferrous Hb with a quantum yield (Phi) of 0.17, that is higher than the method of direct photoexcitation of heme (Phi = 0.006). Electron transfer from a donor molecule to metHb via FMN was completed within 30 ns. Native-PAGE and IEF electrophoresis indicated no chemical modification of the surface of the reduced Hb. Coencapsulation of concentrated Hb solution (35 g/dL) and the FMN/EDTA system in vesicles covered with a phospholipid bilayer membrane (Hb-vesicles, HbV, diameter: 250 nm) facilitated the metHb photoreduction even under aerobic conditions, and the reduced HbV restored the reversible O(2) binding property. A concentrated HbV suspension ([Hb] = 8 g/dL) was sandwiched with two glass plates to form a liquid layer with the thickness of about 10 microm (close to capillary diameter in tissue, 5 microm), and visible light irradiation (221 mW/cm(2)) completed 100% metHb photoreduction within 20 s. The photoreduced FMNH(2) reacted with O(2) to produce H(2)O(2), which was detected by the fluorescence measurement of the reaction of H(2)O(2) and p-nitrophenylacetic acid. However, the amount of H(2)O(2) generated during the photoreduction of HbV was significantly reduced in comparison with the homogeneous Hb solution, indicating that the photoreduced FMNH(2) was effectively consumed during the metHb reduction in a highly concentrated condition inside the HbV nanoparticles.  相似文献   

11.
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   

12.
We studied the effect of noradrenaline on the methaemoglobin (metHb) concentration in rainbow trout red cells. The erythrocytes were incubated in physiological medium with or without noradrenaline and the percentage of metHb of total Hb content was measured. Noradrenaline lowered the metHb content significantly as compared to controls. To study if the effect of noradrenaline was caused by adrenergic intracellular alkalinization, cells were treated with noradrenaline + carbonic anhydrase or noradrenaline + acetazolamide. Carbonic anhydrase inhibits the adrenergic increase in intracellular pH, but did not reduce the effect of noradrenaline on the metHb concentration. Acetazolamide accentuates the increase in intracellular pH. However, there was no difference in the methaemoglobin content of noradrenaline-incubated and noradrenaline + acetazolamide-incubated cells. These results show that the effect of noradrenaline on the methaemoglobin content is independent from the adrenergic increase in intracellular pH. However, amiloride treatment inhibited the effect of noradrenaline on the methaemoglobin content, suggesting that the protein mediating sodium/proton exchange may also be involved in controlling cellular methaemoglobin levels.  相似文献   

13.
S C Larson  G W Fisher  N T Ho  T J Shen  C Ho 《Biochemistry》1999,38(29):9549-9555
Three recombinant mutants of human fetal hemoglobin (Hb F) have been constructed to determine what effects specific amino acid residues in the gamma chain have on the biophysical and biochemical properties of the native protein molecule. Target residues in these recombinant fetal hemoglobins were replaced with the corresponding amino acids in the beta chain of human normal adult hemoglobin (Hb A). The recombinant mutants of Hb F included rHb F (gamma 112Thr --> Cys), rHb F (gamma 130Trp --> Tyr), and rHb F (gamma 112Thr --> Cys/gamma 130Trp --> Tyr). Specifically, the importance of gamma 112Thr and gamma 130Trp to the stability of Hb F against alkaline denaturation and in the interaction with sickle cell hemoglobin (Hb S) was investigated. Contrary to expectations, these rHbs were found to be as stable against alkaline denaturation as Hb F, suggesting that the amino acid residues mentioned above are not responsible for the stability of Hb F against the alkaline denaturation as compared to that of Hb A. Sub-zero isoelectric focusing (IEF) was employed to investigate the extent of hybrid formation in equilibrium mixtures of Hb S with these hemoglobins and with several other hemoglobins in the carbon monoxy form. Equimolar mixtures of Hb A and Hb S and of Hb A(2) and Hb S indicate that 48-49% of the Hb exists as the hybrid tetramer, which is in agreement with the expected binomial distribution. Similar mixtures of Hb F and Hb S contain only 44% hybrid tetramer. The results for two of our recombinant mutants of Hb F were identical to the results for mixtures of Hb F and Hb S, while the other mutant, rHb F (gamma 130Trp --> Tyr), produced 42% hybrid tetramer. The sub-zero IEF technique discussed here is more convenient than room-temperature IEF techniques, which require Hb mixtures in the deoxy state. These recombinant mutants of Hb F were further characterized by equilibrium oxygen binding studies, which indicated no significant differences from Hb F. While these mutants of Hb F did not have tetramer-dimer dissociation properties significantly altered from those of Hb F, future mutants of Hb F may yet prove useful to the development of a gene therapy for the treatment of patients with sickle cell anemia.  相似文献   

14.
Among the four types of hemoglobin (Hb) M with a substitution of a tyrosine (Tyr) for either the proximal (F8) or distal (E7) histidine in the α or β subunits, only Hb M Saskatoon (βE7Tyr) assumes a hexacoordinate structure and its abnormal subunits can be reduced readily by methemoglobin (metHb) reductase. This is distinct from the other three M Hbs. To gain new insight into the cause of the difference, we examined the ionization states of E7 and F8 Tyrs by UV resonance Raman (RR) spectroscopy and Fe–O(Tyr) bonding by visible RR spectroscopy. Hb M Iwate (αF8Tyr), Hb M Boston (αE7Tyr), and Hb M Hyde Park (βF8Tyr) exhibited two extra UV RR bands at 1,603 cm−1 (Y8a′) and 1,167 cm−1 (Y9a′) arising from deprotonated (ionized) Tyr, but Hb M Saskatoon displayed the UV RR bands of protonated (unionized) Tyr at 1,620 and 1,175 cm−1 in addition to those of deprotonated Tyr. Evidence for the bonding of both ionization states of Tyr to the heme in Hb M Saskatoon was provided by visible RR spectroscopy. These results indicate that βE7Tyr of Hb M Saskatoon is in equilibrium between protonated and deprotonated forms, which is responsible for facile reducibility. Comparison of the UV RR spectral features of metHb M with that of metHb A has revealed that metHb M Saskatoon and metHb M Hyde Park are in the R (relaxed) structure, similar to that of metHb A, whereas metHb M Iwate, metHb M Boston and metHb M Milwaukee are in the T (tense) quaternary structure.  相似文献   

15.
Reactions of human oxyhemoglobin A with iron(II) compounds have been investigated. Human oxyhemoglobin (HbO2) reacts with aquopentacyanoferrate(II), Fe(II)(CN)5H2O3-, to yield hydrogen peroxide, aquomethemoglobin and Fe(III)(CN)5H2O2-. The reaction follows a second order rate law, first order in the pentacyanide and in HbO2. Since reaction rates are lower in the presence of catalase, the H2O2 produced must promote metHb formation in reactions independent of pentacyanide. Changes in concentrations of effectors (e.g. H+, inositol hexaphosphate, Cl-, and Zn2+), alkylation of beta-93 cysteine with N-ethylmaleimide, and substitution at distal histidine (as in Hb Zurich with beta-63 His----Arg) in each case can markedly affect pentacyanide reaction rates demonstrating a fine control of rates by protein structure. Hexacyanoferrate(II) (ferrocyanide) reacts with HbO2 to produce cyano-metHb as well as aquo-metHb but the reaction with the hexacyanide is much slower than with the aquopentacyanide. Iron(II) EDTA converts HbO2 to deoxy-Hb with no evidence for formation of metHb as an intermediate. These findings support a mechanism in which the pentacyanide anion reacts directly with coordinated dioxygen. One-electron transfers to O2 from both pentacyanide iron(II) and heme iron(II) result in the formation of a mu-peroxo intermediate, HbFe(III)-O-O-Fe(III) (CN)5(3-). Hydrolysis of this intermediate yields metHb . H2O, H2O2, and FeIII(CN)5H2O2-. The reaction of HbO2 with Fe(CN)6(4-) must follow an outer sphere electron transfer mechanism. However, the very slow rate that is seen with Fe(CN)6(4-) could arise entirely from the pentacyanide produced from loss of one cyanide ligand from the hexacyanide. Fe(II)EDTA reacts rapidly with free O2 in solution but can not interact directly with the heme-bound O2 of HbAO2. The dynamic character of the O2 binding sites apparently permits access of the Fe2+ of the pentacyanide to coordinated dioxygen but the protein structure is not sufficiently flexible to allow the larger Fe2+EDTA molecule to react with bound O2. It is necessary for maintenance of the oxygen transport function of the red cell for reductants such as the methemoglobin reductase system, glutathione, and ascorbate to be able to reduce metHb to deoxy-Hb. It is also important for these reductants to be unable to donate an electron to HbO2 to yield H2O2 and metHb. Thus, a mechanistic requirement for the delivery of one-electron directly to the dioxygen ligand, if peroxide is to be produced, enables the protein to protect the oxygenated species from those electron donors normally present in the cell by denying these reductants steric access to coordinated O2.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
We compare three separation techniques for preparative purposes, i.e. ion-exchange chromatography on CM-cellulose, conventional isoelectric focusing in polyacrylamide gel slabs and immobilized pH gradients. The biological system used to test the three methods is a solution containing four hemoglobin (Hb) valence intermediates, i.e. metHb, oxyHb, (alpha + beta O2)2 and (alpha O2 beta +)2. The delta pI between the two valence intermediates is 0.04 pH units. Immobilized pH gradients give the best performance in terms of resolving power, total amount of protein which can be loaded and retention of biological activity by the protein (the latter assessed by determination of CO dissociation rates).  相似文献   

17.
The complete primary structure has been determined for an inhibitor protein of protein kinase C. The bovine brain-derived inhibitor has a pI of 6 and its N-terminal alanine residue is blocked by acetylation. Fragments obtained by chemical and enzymatic cleavage of the purified inhibitor were analyzed by Edman degradation, fast atom bombardment mass spectrometry, and tandem mass spectrometry. The results establish that the protein has a calculated average molecular mass of 13,690 daltons and contains 125 amino acid residues with the following sequence: (sequence: see text) The inhibitor does not show significant homology with any other known protein. Circular dichroism of the freshly prepared apoprotein indicated a secondary structural content of 23% alpha-helix, 31% beta-sheet, and 11% beta-turn. Immobilization on nitrocellulose followed by exposure to a 65Zn2(+)-containing overlay solution showed that, like protein kinase C itself, the inhibitor is a zinc-binding protein, although the sequence does not reveal a "zinc finger" structure. Competition with 10-fold molar excess Ca2+ or Mg2+ did not reduce the zinc-binding specificity of this inhibitor.  相似文献   

18.
Glutathionylated hemoglobin (Hb-SSG) is now recognized as a promising biomarker of systemic oxidative stress. Aim of this study is to gain a mechanistic insight into its formation. The ability of GSSG to form Hb-SSG through a thiol-disulfide exchange mechanism was firstly examined. For this purpose, GSSG (ranging from 0.23 to 230 μmol/g Hb, 15 μM–15 mM final concentrations) was incubated with 1 mM Hb and the relative content of Hb-SSG determined by direct infusion mass spectrometry (Orbitrap as analyzer). No detectable Hb-SSG was observed at a GSSG concentration range found in physiopathological conditions (0.13–0.23 μmol/g Hb). To reach a detectable Hb-SSG signal, the GSSG concentration was raised to 2.3 μmol/g Hb (0.5% relative abundance). The relative content of Hb-GSSG dose-dependently increased to 6% and 11% at 77 and 153 μmol/g Hb, respectively. The second step was to demonstrate whether Hb-SSG is formed through a sulfenic acid intermediate, a well-recognized mechanism of S-protein glutathionylation. Cys β93 sulfenic acid was found to be formed by oxidizing Hb with 1 mM H2O2, as demonstrated by direct infusion and LC–ESI-MS/MS experiments and using dimedone as derivatazing agent. When H2O2-treated Hb was incubated with physiological concentrations of GSH (9 μmol/g Hb), the corresponding Hb-SSG form was detected, reaching 15% of relative abundance. In summary, we here demonstrate that Hb glutathionylation can occur through a Cys sulfenic acid intermediate which is formed in oxidizing conditions. Hb glutathionylation is also mediated by a thiol-disulfide transfer mechanism, but this requires a concentration of GSSG which is far to be achieved in physiopathological conditions.  相似文献   

19.
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   

20.
Glutathionylation of hemoglobin (Hb) was studied by incubation of intact human erythrocytes with 1 mM tert-butylhydroperoxide (tBHP). Electrophoresis of the membranes showed a time dependent increase of membrane-bound Hb alpha chain until 10 min, and immunoblotting study showed that membrane-bound Hb alpha chain reacted with anti-glutathione antibody only after 10 min. Concomitant with the Hb alpha chain, membrane associated actin, spectrin, and glyceraldehyde 3-phosphate dehydrogenase reacted with the antibody. Cytosolic Hb of the control erythrocytes reacted with anti-glutathione antibody. Together with our previous paper, the present study indicates that at least three different types of glutathionylation of Hb can exist in erythrocytes. The first type is a mixed disulfide bond between reduced glutathione (GSH) and normal Hb. The second type is a disulfide bond between the cysteine 93 of metHb beta chain and oxidized glutathione (GSSG), and the third type is a disulfide bond between the other cysteine residues of metHb alpha chain and/or metHb beta chain and GSSG.  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