首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
1). During the sol to gel transformation of deoxygenated sickle hemoglobin, a time-dependent process preceding gel formation (lag phase) was demonstrated that was inversely proportional to a function of the hemoglobin concentration and that occurred without alteration in temperature, pH, or oxygen tension. 2). As determined by the Schachman modification of the capillary viscometer, preparations of oxyhemoglobin S and A and deoxyhemoglobin A were indistinguishable when compared over a wide range of concentrations. Up to the concentration at which gelling occurred, deoxyhemoglobin S exhibited the same viscosity behavior. The viscosity of deoxygenated hemoglobin S within the lower gelling concentration range was normal during the lag phase and became abnormally high only at the time of gelation.  相似文献   

2.
Methyl acetyl phosphate binds to the 2,3-diphosphoglycerate (2,3-DPG) binding site of hemoglobin and selectively acetylates three amino groups at or near that site. The subsequent binding of 2,3-DPG is thus impeded. When intact sickle cells are exposed to methyl acetyl phosphate, their abnormally high density under anaerobic conditions is reduced to the density range of oxygenated, nonsickling erythrocytes. This change is probably due to a combination of direct and indirect effects induced by the specific acetylation. The direct effect is on the solubility of deoxyhemoglobin S, which is increased from 17 g/dL for unmodified hemoglobin S to 22 g/dL for acetylated hemoglobin S at pH 6.8. Acetylated hemoglobin S does not gel at pH 7.4, up to a concentration of 32 g/dL. The indirect effect could be due to the decreased binding of 2,3-DPG to deoxyhemoglobin S within the sickle erythrocyte, thus hindering the conversion of oxyhemoglobin S to the gelling form, deoxyhemoglobin S.  相似文献   

3.
Sesamol (3,4-methylenedioxyphenol), a monophenolic antioxidant in sesame iol, produced methemoglobin from hemoglobin A (oxyhemoglobin and deoxyhemoglobin) and from red cells. The activity of the compound was more extensive than the polyphenolic compounds. The profiles of the methemoglobin formation by the compound were compared with those by nitrite and hydroxylamine. The formation of methemoglobin from oxyhemoglobin by the compound was rather slowly progressed, but the amount of methemoglobin formed was proportional to the concentration of oxyhemoglobin even when the concentration of the compound was low. The sesamol-induced methemoglobin formation was influenced by inositol hexaphosphate, an allosteric effector of hemoglobin. Thus, the phosphate enhanced the transformation of oxyhemoglobin and inhibited the transformation of deoxyhemoglobin.  相似文献   

4.
Apohemoglobin S formed a gel in the cold (5 degrees C) with a protein concentration in the supernatants after centrifugation of the gels (Csat) near 27 g/dl, in 0.02 M phosphate buffer at pH 7.2. Under the same experimental conditions in mixtures of apohemoglobin S and deoxyhemoglobin S the solubility of hemoglobin S in the cold was decreased from Csat greater than 40 g/dl in the absence to about 18 g/dl in the presence of apohemoglobin S. Conversely, in the same mixture, Csat of apohemoglobin S was decreased to about 5 g/dl. Also, gelling occurred in mixtures of oxyhemoglobin S and its apoderivative. Apohemoglobin A alone did not form gels; however, it induced fiber formation in deoxyhemoglobin S in the cold; unlike apohemoglobin S, it was not included in the precipitate. Gels of apohemoglobin S were not birefringent, and inspection at the electron microscope failed to show the presence of organized structures. Excluded volume effects were probably at the origin of the decreased solubility of hemoglobin S and apohemoglobin S in the presence of each other.  相似文献   

5.
The normal and differential titration curves of liganded and unliganded hemoglobin were measured at various KCl concentrations (0.1 to 2.0 M). In this range of KCl concentrations, the curves for deoxyhemoglobin showed no salt-induced pK changes of titratable groups. In the same salt concentration range oxyhemoglobin showed a marked change in titration behavior which could only be accounted for by a salt-induced increase in pK of some titratable groups. These results show that the suppression of the alkaline Bohr effect by high concentrations of neutral univalent salt is not caused by a weakening of the salt bridges in deoxyhemoglobin but is due to an interaction of chloride ions with oxyhemoglobin. Measurements of the Bohr effect at various KCl concentrations showed that at low chloride ion concentration (5 times 10-3 M) the alkaline Bohr effect is smaller than at a concentration of 0.1 M. This observation indicates that at a chloride ion concentration of 0.1 M, part of the alkaline Bohr effect is due to an interaction of chloride ions with hemoglobin. Furthermore, at low concentrations of chloride ions the acid Bohr effect has almost vanished. This result suggests that part of the acid Bohr effect arises from an interaction of chloride ions with oxyhemoglobin. The dependence of the Bohr effect upon the chloride ion concentration can be explained by assuming specific binding of chloride ions to both oxy- and deoxyhemoglobin, with deoxyhemoglobin having the highest affinity.  相似文献   

6.
The binding of folyl- and antifolylpolyglutamates to hemoglobin   总被引:2,自引:0,他引:2  
A binding method that detects only the strongest binding site for a ligand on a protein has been used to show that folates and folate analogs, conjugated with poly-gamma-glutamates, are bound to hemoglobin. When the concentration of hemoglobin is much larger than that of the polyglutamate, as is the case in the red cell, the fraction bound is a direct function of the hemoglobin concentration and is independent of the total polyglutamate concentration. Binding to deoxyhemoglobin tetramers is competitive with 2,3-diphosphoglycerate. In oxyhemoglobin the folyl and methotrexate polyglutamates are bound preferentially by free alpha beta dimers, but removal of the pteridine moiety leads to tetramer binding even in oxyhemoglobin. Changes in the length of the polyglutamate side chain and alterations of the pteridine structure such as reduction and/or methylation have a much larger effect on the constant for binding to deoxyhemoglobin tetramers than on that for oxyhemoglobin dimers. The implications of these results for the storage of pteroylpolyglutamates in the erythrocyte and their release from the red cell under the influence of the degree of oxygenation and variations in the 2,3-diphosphoglycerate level are discussed.  相似文献   

7.
Quasi-elastic light scattering has been used to measure the change in the translational diffusion coefficient of hemoglobin upon oxygenation and the difference in the diffusion coefficients of oxy- and methemoglobin. The diffusion coefficients of oxy- and methemoglobin were found to be the same within the experimental accuracy of 0.2%, while the diffusion coefficient of oxyhemoglobin tetramers in solution at 13 mg/ml was found to be 0.8% smaller than that of deoxyhemoglobin at the same concentration, when the reversible dissociation of oxyhemoglobin tetramers into dimers was taken into account. In the limit of zero concentration, the oxyhemoglobin diffusion coefficient was found to be 1.5% ± 1.0% smaller than that of deoxyhemoglobin. This result is in very good agreement with what we predict using atomic coordinates to model the liganded and unliganded hemoglobin molecules as ellipsoids of revolution.  相似文献   

8.
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   

9.
Under physiological conditions of pH (7.4) and chloride concentration (0.15 M), the oxygen affinity of bovine hemoglobin is substantially lower than that of human hemoglobin. Also, the Bohr effect is much more pronounced in bovine hemoglobin. Numerical simulations indicate that both phenomena can be explained by a larger preferential binding of chloride ions to deoxyhemoglobin in the bovine system. Also, they show that the larger preferential binding may be produced by a decreased affinity of the anions for oxyhemoglobin, thereby stressing the potential relevance of the oxy conformation in regulating the functional properties of the protein. The conformation of the amino-terminal end of the beta subunits appears to regulate the interaction of hemoglobin with solvent components. The pronounced sensitivity of the oxygen affinity of bovine hemoglobin to chloride concentration and to pH suggests that in bovine species these are the modulators of oxygen transport in vivo.  相似文献   

10.
Human hemoglobin was spin labeled with 4-isothiocanato-2,2,6,6-tetramethyl-piperdinooxyl, which is known to bind specifically to the N-terminal alpha-amino groups of proteins and slightly to the reactive sulfhydryl groups. Electron spin resonance (ESR) analysis indicated a partially resolved five-line spectrum, suggesting that the label was attached to at least two different binding sites. Using specific blocking reagents prior to spin labeling, the two binding sites were attributed to the sulfhydryl group of beta-93 (immobile) and the alpha-amino group of the N-terminal valines (mobile). The relative motion of the spin at one set of binding sites was restricted regardless of the state of ligation and pH, while the motion at the other site showed dependence on those parameters, e.g. the spin-labeled N-terminal ends of deoxyhemoglobin have restricted motion at all pH ranges studied, while those of oxyhemoglobin are relatively free to move at the basic pH range, but become more restricted in the acidic pH range.  相似文献   

11.
The physical properties of deoxyhemoglobin S gels formed from solutions at concentrations and temperatures approaching those in vivo have been characterized by stress relaxation using a rotational rheometer. Gels were annealed in the rheometer and then subjected to a constant shear strain; thereafter the stress sustained was followed with time. Gels with solid-like behavior held stress indefinitely, and were characterized by yield temperature (the temperature at which stress decreased). Gels with less solid behavior were unable to hold target stress, and were characterized by yield stress (maximum stress attained) and equilibrium stress (final stress held). The samples were ultracentrifuged to calculate pellet and polymer masses. The solidity of the gels, as measured by yield temperature or yield stress, was related to the initial hemoglobin concentration, pellet and polymer masses, shear history, temperature, and the temperature and time of annealing. Solidity increased significantly with time when gels were annealed at 37 degrees C, whereas, when annealed at 25 degrees C, no or minimal increases in solidity were noted. Studies suggest that polymerization occurs rapidly and is completed early in or before the gel annealing period and that the increase in solidity with time of annealing is mainly due to factors other than polymer mass, i.e. alignment, increasing bond strength, water loss. The chemical activity of deoxyhemoglobin S did not affect the solidity of the formed gels. When the resultant polymer masses were comparable, gels formed from samples with albumin present (higher initial total protein concentration, but lower initial deoxyhemoglobin S concentration), had the same behavior as gels formed from solutions with higher initial hemoglobin S concentration. These findings demonstrate that gel annealing conditions must be standardized when comparing the rheologic behaviors of deoxyhemoglobin S gels and indicate that the gel's physical properties (influenced by polymer mass, shear history, annealing time) must be considered in understanding pathophysiology of sickling disorders.  相似文献   

12.
Collective motion of iron and exogenous ligands in hemoglobin   总被引:1,自引:0,他引:1  
A Levy 《Biochemistry》1989,28(18):7144-7147
M?ssbauer studies of deoxyhemoglobin (Hb), oxyhemoglobin (HbO2), and (carbonmonoxy)-hemoglobin (HbCO) have been performed providing information regarding the center shift (CS) up to 250 K and the recoil-free fraction (f) in the "solid state regime". The temperature dependence of the CS indicates that the effective mass of the iron in liganded hemoglobin includes the exogenous ligand, suggesting that vibrations of the iron are strongly coupled to those of the ligand. As a result of this coupled motion, the recoil-free fraction (f) provides a valuable probe for the shape of the potential energy surface in the ligand pocket. Thus HbO2 experiences a stronger restoring force than HbCO, which can be attributed to a steeper potential well for the oxygen into the ligand pocket.  相似文献   

13.
Abstract

Tritium labeled asparagine binds to oxyhemoglobin S and to a mixture of hemoglobins C and S in the molar ratio of 3.38:1 and 8.2:1 respectively. From the dialysis equilibrium studies it appears that labeled asparagine does not bind to oxy- or deoxy- hemoglobin A nor to deoxyhemoglobin S. The constant for equilibrium association of asparagine for oxyhemoglobin S is 7.38 × 107 M?1 and for'oxyhemoglobin CS 4.8 × 104 M?1 at 23°C. Tritium labeled asparagine is bound to oxyhemoglobin S and CS sufficiently strongly to prevent dissociation under the conditions of gel electrophoresis at pH 9.50. The protein with and without bound asparagine, gluta-mine or homoserine, is indistinguishable in molecular net charge and size by the criteria of quantitative polyacrylamide gel electrophoresis (PAGE). Also there were no significant differences in mobility between hemoglobin S and hemoglobin C in the presence and absence of asparagine, glutamine and homoserine as detectable in agar coated cellulose acetate electrophoresis at pH 6.3. Erythrocytes containing hemoglobin S and CS, after incubation with tritium labeled asparagine and lysis under the conditions of gel electrophoresis at pH 9.5, release hemoglobin S and C with bound tritiated asparagine. No tritiated asparagine remains bound to the ghost.  相似文献   

14.
Equations for the spectrophotometric analysis of hemoglobin mixtures   总被引:11,自引:0,他引:11  
The extinction coefficients for deoxyhemoglobin, oxyhemoglobin, and carbonmonoxyhemoglobin as well as those of ferrihemoglobin between pH 6.2 and 8.8 are given at a number of wavelengths. Equations are presented for the analysis of ternary mixtures of these hemoglobin derivatives.  相似文献   

15.
J B Matthew  G I Hanania  F R Gurd 《Biochemistry》1979,18(10):1919-1928
The modified Tanford-Kirkwood theory of Shire et al. [Shire, S. J., Hanania, G.I.H., & Gurd, F.R.N. (1974) Biochemistry 13, 2967] for electrostatic interactions was applied to the hydrogen ion equilibria of human deoxyhemoglobin and oxyhemoglobin. Atomic coordinates for oxyhemoglobin were generated by the application of the appropriate rigid rotation function to alpha and beta chains of the deoxyhemoglobin structure [Fermi, G. (1975) J. Mol. Biol. 97, 237]. The model employs two sets of parameters derived from the crystalline protein structures, the atomic coordinates of charged amino acid residues and static solvent accessibility factors to reflect their individual degrees of exposure to solvent. Theoretical titration curves based on a consistent set of pKint values compared closely with experimental potentiometric curves. Theoretical pK values at half-titration for individual protein sites corresponded to available observed values for both quaternary states. The results bring out the cumulative effects of numerous electrostatic interactions in the tetrameric structures and the major effects of the quaternary transition that result from changes in static solvent accessibility of certain ionizable groups.  相似文献   

16.
Hydroxyurea represents an approved treatment for sickle cell anemia and acts as a nitric oxide donor under oxidative conditions in vitro. Electron paramagnetic resonance spectroscopy shows that hydroxyurea reacts with oxy-, deoxy-, and methemoglobin to produce 2-6% of iron nitrosyl hemoglobin. No S-nitrosohemoglobin forms during these reactions. Cyanide and carbon monoxide trapping studies reveal that hydroxyurea oxidizes deoxyhemoglobin to methemoglobin and reduces methemoglobin to deoxyhemoglobin. Similar experiments reveal that iron nitrosyl hemoglobin formation specifically occurs during the reaction of hydroxyurea and methemoglobin. Experiments with hydroxyurea analogues indicate that nitric oxide transfer requires an unsubstituted acylhydroxylamine group and that the reactions of hydroxyurea and deoxy- and methemoglobin likely proceed by inner-sphere mechanisms. The formation of nitrate during the reaction of hydroxyurea and oxyhemoglobin and the lack of nitrous oxide production in these reactions suggest the intermediacy of nitric oxide as opposed to its redox form nitroxyl. A mechanistic model that includes a redox cycle between deoxyhemoglobin and methemoglobin has been forwarded to explain these results that define the reactivity of hydroxyurea and hemoglobin. These direct nitric oxide producing reactions of hydroxyurea and hemoglobin may contribute to the overall pathophysiological properties of this drug.  相似文献   

17.
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   

18.
The transverse water proton relaxation times (T2) of erythrocytes homozygous and heterozygous for hemoglobin S have been measured as a function of oxyhemoglobin concentration at 37 °C. An immediate decrease in T2 is observed in S/S erythrocytes as the amount of oxyhemoglobin is decreased and the maximum change is observed at 50% deoxyhemoglobin S. In heterozygous erythrocytes, the T2 remains unchanged until a critical level of deoxyhemoglobin is attained. The critical level of deoxyhemoglobin is a function of the percentage of hemoglobin S in the heterozygous erythrocytes. A Hill plot of the data obtained from S/S erythrocytes gives an n value of around 2.4. These results suggest that the measurement of T2 is sensitive to the very early stages of the polymerization process. This suggestion is supported by calculations; our T2 measurements are sensitive to a range of correlation times expected for hemoglobin monomers at one extreme and linear polymers of seven hemoglobin molecules at the other extreme.  相似文献   

19.
Quasi-elastic light scattering has been used to examine solutions and gels of deoxyhemoglobin S. The autocorrelation function is found to decay with a characteristic exponential relaxation which can be ascribed to the diffusion of monomer (64,000 molecular weight) hemoglobin S molecules. In the absence of polymers, the relaxation time is in good agreement with previous measurements of the diffusion coefficient for solutions of normal human hemoglobin. In the presence of the polymer phase, a large (greater than 200-fold) increase in the scattered intensity is observed but no contribution to the decay of the autocorrelation function from the motion of the aligned polymer phase can be detected. Heterodyning between the time-independent scattering amplitude from the polymers and the time-dependent scattering of the diffusing monomers results in a twofold increase in the relaxation time arising from monomer diffusion.  相似文献   

20.
The contribution of the interaction of chloride ions with deoxy and oxyhemoglobin to the Bohr effect can be described by a simple binding model. Applying this model to experiment data reveals that at physiological pH and ionic strength about half of the release of Bohr protons is due to a difference in chloride ion binding to deoxy- and oxyhemoglobin. The chloride-independent part of the Bohr effect corresponds with the shift in pK which His-146 beta shows upon oxygenation. The proton absorptioon by hemoglobin observed upon oxygenation below pH 6 is apparently due to a chloride-ion-induced proton uptake, which is larger for oxyhemoglobin than for deoxyhemoglobin. The analysis of the experimental data indicates the existence of only two oxygen-linked chloride ion binding sites in both deoxy and oxyhemoglobin. In deoxyhemoglobin the binding sites most likely consist of Val-1 alpha of one chain and Arg-141 alpha of the partner chain. The sites in oxyhemoglobin consist of groups with a pK value in the neutral pH range; they do not contain lysyl or arginyl residues.  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