首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Peroxynitrite-mediated oxidative chemistry is currently the subject of intense investigation owing to the toxic side effects associated with nitric oxide overproduction. Using direct electron spin resonance spectroscopy (ESR) at 37 degrees C, we observed that in human erythrocytes peroxynitrite induced a long-lived singlet signal at g = 2.004 arising from hemoglobin. This signal was detectable in oxygenated red blood cells and in purified oxyhemoglobin but significantly decreased after deoxygenation. The formation of the g = 2.004 radical required the presence of CO2 and pH values higher than the pKa of peroxynitrous acid (pKa = 6.8), indicating the involvement of a secondary oxidant formed in the interaction of ONOO- with CO2. The g = 2.004 radical yield leveled off at a 1:1 ratio between peroxynitrite and oxyhemoglobin, while CO-hemoglobin formed less radical and methemoglobin did not form the radical at all. These results suggest that the actual oxidant is or is derived from the ONOOCO2- adduct interacting with oxygenated FeII-heme. Spin trapping with 2-methyl-2-nitrosopropane (MNP) of the g = 2.004 radical and subsequent proteolytic digestion of the MNP/hemoglobin adduct revealed the trapping of a tyrosyl-centered radical(s). A similar long-lived unresolved g = 2.004 singlet signal is a common feature of methemoglobin/H2O2 and metmyoglobin/H2O2 systems. We show by spin trapping that these g = 2.004 signals generated by H2O2 also indicated trapping of radicals centered on tyrosine residues. Analysis of visible spectra of hemoglobin treated with peroxynitrite revealed that, in the presence of CO2, oxyhemoglobin was oxidized to a ferryl species, which rapidly decayed to lower iron oxidation states. The g = 2.004 radical may be an intermediate formed during ferrylhemoglobin decay. Our results describe a new pathway of peroxynitrite-dependent hemoglobin oxidation of dominating importance in CO2-containing biological systems and identify the g = 2.004 radical(s) formed in the process as tyrosyl radical(s).  相似文献   

2.
O2 transport was examined by measuring the fractional saturation of concentrated hemoglobin solutions flowing through an artificial capillary that was approximately 27 micron in diameter and embedded in a silicone rubber film approximately 170 micron thick. The effects of pH, hemoglobin concentration, O2 tension, temperature, and organic phosphate were measured and analyzed quantitatively by a rigorous mathematical model that included the geometry of the capillary in the silicone film, parabolic flow velocity distributions inside the lumen, and cooperative O2 binding by hemoglobin. The rates of both oxygenation and deoxygenation were limited by diffusion and governed by the magnitude of the O2 gradient between the intracapillary fluid phase and the external gas space. In uptake experiments, O2 flux is determined primarily by the external O2 tension (16-160 mmHg in our experiments) because the internal O2 pressure is kept small due to chemical combination with hemoglobin. In release experiments, the external O2 tension is maintained at zero, and the transport rate is determined by the intracapillary partial pressure of O2 that is proportional to the O2 half-saturation pressure of hemoglobin value of the hemoglobin sample. As a result, factors that change the affinity of hemoglobin for O2, such as pH, temperature, and organic phosphate concentration, influence strongly the rate of O2 release but have little effect on the rate of O2 uptake. These properties are physiologically advantageous, since a decrease in pH or an increase in temperature during exercise increases both the rate and extent of deoxygenation while not altering the kinetics of oxygenation.  相似文献   

3.
M J Rao  A S Acharya 《Biochemistry》1992,31(32):7231-7236
Glu-43(beta) of hemoglobin A exhibits a high degree of chemical reactivity around neutral pH for amidation with nucleophiles in the presence of carbodiimide. Such a reactivity is unusual for the side-chain carboxyl groups of proteins. In addition, the reactivity of Glu-43(beta) is also sensitive to the ligation state of the protein [Rao, M. J., & Acharya, A. S. (1991) J. Protein Chem. 10, 129-138]. The influence of deoxygenation of hemoglobin A on the chemical reactivity of the gamma-carboxyl group of Glu-43(beta) has now been investigated as a function of pH (from 5.5 to 7.5). The chemical reactivity of Glu-43(beta) for amidation increases upon deoxygenation only when the modification reaction is carried out above pH 6.0. The pH-chemical reactivity profile of the amidation of hemoglobin A in the deoxy conformation reflects an apparent pKa of 7.0 for the gamma-carboxyl group of Glu-43(beta). This pKa is considerably higher than the pKa of 6.35 for the oxy conformation. The deoxy conformational transition mediated increase in the pKa of the gamma-carboxyl group of Glu-43(beta) implicates this carboxyl group as an alkaline Bohr group. The amidated derivative of hemoglobin A with 2 mol of glycine ethyl ester covalently bound to the protein was isolated by CM-cellulose chromatography.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Kinetics of blood deoxygenation was studied during acute hypoxia induced by subcutaneous administration of sodium nitrite using polarographic method. Plasma oxygen tension remained unaltered as the dose of sodium nitrite increased, while the dynamics of oxygen release was dose-dependent. The constant of oxyhemoglobin deoxygenation rate proved to vary with blood deoxygenation. The nitrite-induced deceleration of oxyhemoglobin deoxygenation was due to the inactivation of a fraction of hemoglobin as well as to the increased hemoglobin oxygen affinity and possible changes in the oxygen permeability of erythrocyte membranes during acute methemoglobinemia.  相似文献   

5.
A small stopped-flow cuvette was built into a computer-controlled Cary 210 spectrophotometer. The enzymatic depletion of oxygen in solutions of hemoglobin and myoglobin was initiated by flowing the hemeproteins with the enzyme against a solution of the hemeproteins containing the appropriate substrate. The deoxygenation was homogeneous throughout the solution. Oxygen activity was calculated at each instant of time from the fractional saturation of Mb, determined from observations at the Hb/HbO2 isosbestic wavelength. Fractional saturation of Hb was determined from absorbances at the Mb/MbO2 isosbestic wavelength. The spectrophotometer cycled between these two wavelengths during the deoxygenation. The deoxygenation of HbO2 was largely complete in 20-25 min, whereas the deoxygenation of MbO2 was allowed to proceed for about 1 h. This procedure eliminates equilibration of Hb solutions with a gas phase and replaces oxygen electrode readings with spectrophotometric sensing by Mb, providing essentially instantaneous determinations of oxygen activity and hence 250-500 or more independent data points per run. The Mb and Hb data vectors require several manipulations to correct for small relative displacements in time and for small non-isosbestic effects. Detailed consideration of the enzyme kinetics allowed oxygen activities to be determined in regions where Mb is a poor sensor. Studies of HbO2 deoxygenation as a function of wavelength show that the determination of the four Adair constants requires in addition the determination of three spectroscopic parameters. Values of the apparent Adair constants, determined without these spectroscopic parameters, depend strongly on the monitoring wavelength.  相似文献   

6.
The effect of dimethyl adipimidate, a bifunctional imidoester, on the oxygen affinity of hemoglobin A has been studied. Treatment of human oxyhemoglobin with 5 mM dimethyl adipimidate at pH 8.5, room temperature is accompanied by an increase in oxygen affinity in the presence and absence of 2,3-diphosphoglyceric acid. Circular dichroism measurements in the ultraviolet region indicate that dimethyl adipimidate-treated hemoglobin exhibits a reduced conformational change upon deoxygenation. In order to study the contribution of reacted individual subunits, alpha and beta subunits of dimethyl adipimidate-treated and untreated hemoglobin have been separated and reconstituted to form hybrid tetramers containing either the alpha-treated (alpha t beta c) or the beta-treated subunits (alpha c beta t). Electrophoresis on sodium dodecyl sulfate polyacrylamide gels of isolated alpha and beta globin subunits as well as hybrid tetramers from dimethyl adipimidate-treated hemoglobin reveals that 20% of the globin subunits are cross-linked. In the absence of 2,3-diphosphoglyceric acid, modification of alpha subunits increases the oxygen affinity and reduces the conformational change of the tetramer upon deoxygenation whereas modification of beta subunits has no effect. However, treatment of beta subunits decreases the effect of 2,3-diphosphoglyceric acid on the oxygen affinity of the hybrids and reduces the 2,3-diphosphoglyceric acid-induced spectral changes in oxyhemoglobin. Therefore the interaction of dimethyl adipimidate with both the alpha and beta subunits contributes to regulating the oxygen affinity of human hemoglobin.  相似文献   

7.
The changes in cerebral metabolism in mice in severe hypoxia were investigated by analyses of changes in the levels of energy metabolites and near-infrared spectrophotometric assessment of the states of hemoglobin and cytochrome oxidase. Under 4.4% O2, the contribution of anaerobic ATP production was at most about 20% of the demand. However, the cerebral ATP level was kept at the control level until about 1 min before death. Pentobarbital anesthesia, which reduced the cerebral rate of metabolism, prolonged the survival time, although anaerobic ATP production still did not support ATP demand. Under these conditions, deoxygenation of hemoglobin and reduction of cytochrome oxidase proceeded rapidly within 1 min. Hemoglobin reached the maximum state of deoxygenation in the middle phase of hypoxia, with no further change until death. However, cytochrome oxidase was reduced slowly with one phase of partial reoxidation due to increase of cerebral blood volume, and reached the completely reduced state at death. From these results it was concluded that the aerobic ATP synthesis, which supplied more than 80% of the cerebral demand, decreased gradually because of limitation of oxygen supply, and that the failure of oxidative phosphorylation to meet demand triggered the decrease in the cellular ATP level that led to death.  相似文献   

8.
Hemoglobin Fannin-Lubbock was found in a 9-year-old Mexican-American female. The abnormal hemoglobin was detected as a fast-moving variant by electrophoresis on cellulose acetate at pH 8.4. Structural analysis indicated a substitution in the beta-chain of aspartic acid for glycine at position 119, a position involved in the alpha1beta1 contact of the hemoglobin tetramer. This contact between unlike chains is larger and undergoes a smaller shift during the process of oxygenation and deoxygenation that the alpha1beta2 contact (Perutz, M.F., Muirhead, H., Cox, J.M. and Goaman, L.C.G. (1968) Nature 219, 131-139). Mutations in this contact tend to cause slight or no changes in functional behavior. Apart from a mild anemia, the propositus did not exhibit any obvious clinical symptoms.  相似文献   

9.
This paper presents data on the effect of phosphates and polyphosphates on the formation of hemoglobin S fiber, and on the Bohr effect of hemoglobin S samples whose concentration was high enough (near 5 mM) in order to form fibers upon deoxygenation. The experiments were performed in 0.2 M Bistris or Tris buffers at 30 degrees C in the presence and absence of inositol hexakisphosphate and of 2,3-diphosphoglycerate. Alternatively, 0.2 M phosphate buffers were used without addition of effectors. Under these conditions, few fibers were formed in Tris or Bistris buffers, while extensive fiber formation occurred in the presence of phosphates and polyphosphates. In all cases, increasing pH strongly inhibited fiber formation. At pH 7.5 and above, fibers were not formed in our samples. In the presence of phosphates and polyphosphates fiber formation reduced the oxygen affinity of hemoglobin S with respect to either hemoglobin A or soluble hemoglobin S under similar experimental conditions. The fiber-polyphosphate complexes showed a larger Bohr effect than that in hemoglobin A. In the presence of inositol hexakisphosphate fiber-forming solutions of hemoglobin S liberated as much as six protons per tetramer upon oxygen binding. The increased liberation of protons was probably due to a higher affinity of the effectors for the fibers of hemoglobin S. Very likely the higher affinity was supported by a conformational change of hemoglobin S specific for the fibers.  相似文献   

10.
Changes of osmolality (Osm) were measured by freezing-point determination in true plasma of 10 healthy subjects. This was done after equilibration with CO2 (0.5-10.0%), after the addition of lactic acid (10 and 20 mmol/l), and after deoxygenation. The graph for the dependence of Osm on CO2 partial pressure (PCO2) in oxygenated blood resembles the classical CO2 absorption curve. The increase of Osm with PCO2 (approximately 0.2 mosmol . kg H2O-1 . Torr-1) is almost as great as the increase in dissolved CO2 plus bicarbonate (HCO-3). Addition of lactic acid shifts the curve upward by only 0.6 mosmol/mmol because of displacement of HCO-3. Deoxygenation has no significant effect at constant PCO2 despite an increase in [HCO-3]. This is probably due to the binding of 2,3-diphosphoglycerate to hemoglobin. It can be seen in the Osm-pH diagram that differences between CO2 and lactic acid titration largely disappear. For each lactic acid concentration there is a linear dependence corresponding to the linear [HCO-3]-pH relation in plasma. At constant pH, Osm increases after deoxygenation. The observed in vitro relation might explain part of the osmolality increase during physical exercise.  相似文献   

11.
The monomer-dimer equilibrium and the oxygen binding properties of ferrous recombinant Vitreoscilla hemoglobin (Vitreoscilla Hb) have been investigated. Sedimentation equilibrium data indicate that the ferrous deoxygenated and carbonylated derivatives display low values of equilibrium dimerization constants, 6 x 10(2) and 1 x 10(2) M(-1), respectively, at pH 7.0 and 10 degrees C. The behavior of the oxygenated species, as measured in sedimentation velocity experiments, is superimposable to that of the carbonylated derivative. The kinetics of O(2) combination, measured by laser photolysis at pH 7.0 and 20 degrees C, is characterized by a second-order rate constant of 2 x 10(8) M(-1) s(-1) whereas the kinetics of O(2) release at pH 7.0 is biphasic between 10 and 40 degrees C, becoming essentially monophasic below 10 degrees C. Values of the first-order rate constants (at 20 degrees C) and of the activation energies for the fast and slow phases of the Vitreoscilla Hb deoxygenation process are 4.2 s(-1) and 19.2 kcal mol(-1) and 0.15 s(-1) and 24.8 kcal mol(-1), respectively. Thus the biphasic kinetics of Vitreoscilla Hb deoxygenation is unrelated to the association state of the protein. The observed biphasic oxygen release may be accounted for by the presence of two different conformers in thermal equilibrium within the monomer. The two conformers may be assigned to a structure in which the heme-iron-bound ligand is stabilized by direct hydrogen bonding to TyrB10 and a structure in which such interaction is absent. The slow interconversion between the two conformers may reflect a very large conformational rearrangement in the disordered distal pocket segment connecting helices C and E.  相似文献   

12.
Human hemoglobin containing cobalt protoporphyrin IX or cobalt hemoglobin has been separated into two functionally active alpha and beta subunits using a new method of subunit separation, in which the -SH groups of the isolated subunits were successfully regenerated by treatment with dithiothreitol in the presence of catalase. Oxygen equilibria of the isolated subunit chains were examined over a wide range of temperature using Imai's polarographic method (Imai, K., Morimoto, H., Kotani, M., Watari, H., and Kuroda, M. (1970) Biochim. Biophys. Acta 200, 189-196). Kinetic properties of their reversible oxygenation were investigated by the temperature jump relaxation method at 16 degrees. Electron paramagnetic resonance characteristics of the molecules in both deoxy and oxy states were studies at 77K. The oxygen affinity of the individual regenerated chains was higher than that of the tetrameric cobalt hemoglobin and was independent of pH. The enthalpy changes of the oxygenation have been determined as -13.8 kcal/mol and -16.8 kcal/mol for the alpha and beta chains, respectively. The rates of oxygenation were similar to those reported for iron hemoglobin chains, whereas those of deoxygenation were about 10(2) times larger. The effects of metal substitution on oxygenation properties of the isolated chains were correlated with the results obtained previously on cobalt hemoglobin and cobalt myoglobin. The EPR spectrum of the oxy alpha chain showed a distinctly narrowed hyperfine structure in comparison with that of the oxy beta chain, indicating that the environment around the paramagnetic center (the bound oxygen) is different between these chains. In the deoxy form, EPR spectra of alpha and beta chains were indistinguishable. These observations suggest that one of the inequivalences between alpha and beta chains might exist near the distal histidine group.  相似文献   

13.
The 3-fold increase in the carbamylation rate of Val-1 (alpha) of hemoglobin upon deoxygenation described earlier is now shown to be a sensitive probe of conformational change. Thus, whereas this residue in methemoglobin A is carbamylated at the same rate as in liganded hemoglobin, upon addition of inositol hexaphosphate its carbamylation rate is enhanced 30% as much as the total change in the rate between the CO and deoxy states. For CO-hemoglobin Kansas in the presence of the organic phosphate, the relative increase in the carbamylation rate of this residue is about 50%. These results indicate that methemoglobin A and hemoglobin Kansas in the presence of inositol hexaphosphate do not assume a conformation identical with deoxyhemoglobin but rather form either a mixture of R and T states or an intermediate conformation in the region around Val-1 (alpha). Studies on the mechanism for the rate enhancement in deoxyhemoglobin suggest that the cyanate anion binds to groups in the vicinity of Val-1 (alpha) prior to proton transfer and carbamylation of this NH2-terminal residue. Thus, specific removal with carboxypeptidase B of Arg-141 (alpha), which is close to Val-1 (alpha) in deoxyhemoglobin, abolishes the enhancement in carbamylation. Chloride, which has the same valency as cyanate, is a better competitive inhibitor of the carbamylation of deoxyhemoglobin (Ki = 50 mM) compared with liganded hemoglobin. Nitrate and iodide are also effective inhibitors of the carbamylation of Val-1 (alpha) of deoxyhemoglobin (Ki = 35 mM); inorganic phosphate, sulfate, and fluoride are poor competitive inhibitors. The change in pKa of Val-1 (alpha) upon deoxygenation may be due to its differential interaction with chloride.  相似文献   

14.
The deoxygenation kinetics of isolated adult and fetal hemoglobin are measured. The results demonstrate that significant functional differences exist between the two tetrameric hemoglobins. It is pointed out that these functional differences closely parallel the differences in similar properties of beta and gamma chains. It is also shown that 2,3-diphosphoglycerate (2,3-DPG) has no significant effect on the deoxygenation rate of fetal hemoglobin. This result appears to be consistent with the reported weaker binding of 2,3-DPG to the oxygen linked groups of fetal hemoglobin.  相似文献   

15.
1. Spectrophotometric determination of oxygen uptake using oxyhemoglobin as oxygen donor and indicator was used for assay of H2O2-generating oxidases like monoamine oxidase and glucose oxidase. 2. In order to decompose H2O2 formed during the oxygen uptake, catalase and methanol (or ethanol) was added to the respiratory system. At pH values higher than 7.5 the oxydation of deoxygenated hemoglobin to methemoglobin was less than 3%. 2. Oxidases with low Km for oxygen can be assayed using the spectrophotometric method if suitable correction factors are introduced into the calculation of oxygen uptake. The correction factor represents the ratio of the rate of formation (or disappearance) of one of the reactants and the rate of oxyhemoglobin deoxygenation, measured under identical experimental conditions.  相似文献   

16.
The dissociation and oxygen equilibrium properties of whole blood and the purified hemoglobin from Eisenia foetida were compared. Oxygen affinities agreed approximately with each other in the range of pH 6.0 to 9.5. The values of n1/2 were higher in whole blood than in the purified hemoglobin between pH 7.0 and 9.5. The maximum values, obtained near pH 8, were about 6 in whole blood and 3.5 in the purified hemoglobin. In the purified hemoglobin, alkaline dissociation started at pH 7.8, and the approximately 60 S whole molecule dissociated completely into approximately 10 S and 5-6 S components at pH 9.1. In whole blood, however, the dissociation started at pH 8.2 and the complete disappearance of the approximately 60 S molecule occurred at pH 9.6. The values of n1/2 for the dissociation products were lower than those of the purified hemoglobin between pH 7.0 and 9.0. The value of n1/2 decreased with increasing dissociation of the approximately 60 S whole molecule with a pH rise in both whole blood and the purified hemoglobin. Addition of CaCl2 or MgCl2 up to 10 mM to the purified hemoglobin at pH 8.0-8.1 induced increases in oxygen affinity and cooperativity and in the stability of the approximately 60 S whole molecule. The effect on the oxygenation properties was greater with CaCl2 than MgCl2 at the same molar concentration. The stabilizing effect on the approximately 60 S molecule was almost the same with both CaCl2 and MgCl2. These results suggest that the dissociation of property of the hemoglobin in whole blood is controlled by both Ca2+ and Mg2+, and that its oxygenation property is controlled by Ca2+.  相似文献   

17.
The temperature and cell volume dependence of the NMR water proton line-width, spin-lattice, and spin-spin relaxation times have been studied for normal and sickle erythrocytes as well as hemoglobin A and hemoglobin S solutions. Upon deoxygenation, the spin-spin relaxation time (T2) decreases by a factor of 2 for sickle cells and hemoglobin S solutions but remains relatively constant for normal cells and hemoglobin A solutions. The spin-lattice relaxation time (T1) shows no significant change upon deoxygenation for normal or sickle packed red cells. Studies of the change in the NMR linewidth, T1 and T2 as the cell hydration is changed indicate that these parameters are affected only slightly by a 10-20% cell dehydration. This result suggests that the reported 10% cell dehydration observed with sickling is not important in the altered NMR properties. Low temperature studies of the linewidth and T1 for oxy and deoxy hemoglobin A and hemoglobin S solutions suggest that the "bound" water possesses similar properties for all four species. The low temperature linewidth ranges from about 250 Hz at -15 degrees C to 500 Hz at -36 degrees C and analysis of the NMR curves yield hydration values near 0.4 g water/g hemoglobin for all four species. The low temperature T1 data go through a minimum at -35 degrees C for measurements at 44.4 MHz and -50 degrees C for measurements at 17.1 MHz and are similar for oxy and deoxy hemoglobin A and hemoglobin S. These similarities in the low temperature NMR data for oxy and deoxy hemoglobin A and hemoglobin S suggest a hydrophobically driven sickling mechanism. The room temperature and low temperature relaxation time data for normal and sickle cells are interpreted in terms of a three-state model for intracellular water. In the context of this model the relaxation time data imply that type III, or irrotationally bound water, is altered during the sickling process.  相似文献   

18.
Transmission electron microscopy has been used to study intracellular sickle hemoglobin polymer in unfractionated cells from the arterial and venous blood of patients and after external deoxygenation. We detect polymerized hemoglobin in up to 10% of the cells in the venous circulation, especially in cells that are "cigar-shaped" and appear to be irreversibly sickled. We could not see well-defined polymer in mixed arterial samples; nevertheless, we found electron opaque spots, which could be ferritin granules, hemosiderin, or small aggregates of hemoglobin S. However, upon sequential chemical deoxygenation using 1.0% sodium metabisulphite, polymer formation was seen at oxygen saturation values of 75%-85%. Cells that were physically deoxygenated using gas mixtures containing nitrogen-carbon dioxide-oxygen mixtures were found to contain distinct polymers of deoxyhemoglobin S at oxyhemoglobin saturation values of 50%-75%. As deoxygenation increases, we detect short, randomly arranged polymer in a loose network, with occasional long polymers. Upon further deoxygenation, the length and number of polymer forms increased. Between 0% and 50% saturation, most erythrocytes were full of long, parallel, closely packed polymers that tend to align and run parallel to the cell membrane. In both chemical and physically deoxygenated blood samples, cells were seen at 50%-75% oxyhemoglobin saturation that retained their normal biconcave disc shape, although they contained significant amounts of polymer. The structural changes in sickle erythrocytes seen in vitro due to physical or chemical deoxygenation of cells, may reflect in vivo intracellular changes in the sickle cell patient.  相似文献   

19.
The O2 binding properties of sulfhemoglobin were studied. The oxygen tension required for half-saturation of sulfhemoglobin is more than 2 orders of magnitude higher than that for hemoglobin A. The binding of O2 exhibits an alkaline Bohr effect larger than that observed for hemoglobin, yet the Hill number is unity. From the Bohr titration curve, 0.68 proton is released during O2 binding at 0 degrees C. Sulfhemoglobin prepared from carboxypeptidase A-treated hemoglobin has an affinity for O2 which is about the same as that of sulfhemoglobin at the theoretical limit of the Bohr titration curve. Like its carboxypeptidase A-treated hemoglobin precursor, this sulfhemoglobin does not bind O2 cooperatively. Thus, sulfhemoglobin appears to be in a high affinity form at alkaline pH and a low affinity form at acid pH, similar to hemoglobin A. These results demonstrate that the magnitude of the Hill number is not always an indicator of the interaction between oxygen binding and other functions in a hemoglobin.  相似文献   

20.
Carp (Cyprinus carpio) hemoglobin readily autoxidizes in blood smears. Quantification of Soret-band absorbance in individual erythrocytes by means of scanning cytophotometry therefore requires more elaborate methods of preparation of blood samples. Of the fixatives that have been tested, suspension of whole blood in isotonic salt solutions containing glutaraldehyde was most suitable. Glutaraldehyde-fixed red blood cells are totally resistant to hemolysis. In the course of fixation, hemoglobin is transformed to methemoglobin. Spectrophotometry indicated extensive similarities between glutaraldehyde-fixed carp methemoglobin and human methemoglobin. In aqueous solutions, the intensity of the Soret-peak was pH-dependent. The allosteric modifier organic polyphosphate caused an R----T transition, resulting in increased molar extinctions. Dried preparations showed Soret-spectra that were not influenced from either pH or organic polyphosphate concentration of the aqueous suspensions in which the erythrocytes had been stored. The same was true for slide preparations of cyanomethemoglobin, easily derived from methemoglobin on addition of potassium cyanide. In the absence of oxygen fresh blood cells from carp slowly transform their hemoglobin into deoxyhemoglobin. Spectra of the intermediate stages of deoxygenation, Hb4(O2)3, Hb4(O2)2 and Hb4(O2), as well as mixtures of these intermediates, could be monitored.  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