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.
A new turbidimetric method for the direct measurement of the solubility of oxy- and deoxyhemoglobins (Hb) in concentrated phosphate buffer has been established. The principle of the method is the formation of a homogeneous emulsion when hemoglobin is introduced in concentrated phosphate buffer. The solubility of the oxy and deoxy forms of Hb A, Hb S, Hb C, Hb F, and Hb CHarlem (beta 6Glu leads to Val, beta 73Asp leads to Asn) has been studied. The solubility of deoxy-Hb S was the lowest and the solubility curve was broader than those of the other hemoglobins indicating that the aggregates of deoxy Hb S require more water to be dissolved. The solubility of oxy- and deoxyhemoglobins depends on temperature and pH. The solubility of hemoglobins is increased as the temperature is lowered and the pH is raised. The pH dependency of the solubility of deoxy-Hb S in high phosphate buffer was opposite to that of the minimum gelling concentration of deoxy-Hb S. The order of the solubility of Hb CHarlem, Hb FS, Hb AS, Hb CS, and Hb S in concentrated phosphate buffer corresponds to the order of minimum gelling concentration of these hemoglobins or hemoglobin mixtures. Solubility studies of a 1:1 mixture of deoxy-Hb A and deoxy-Hb S show that deoxy-Hb A aggregates in 2.42 M phosphate buffer in which pure deoxy-Hb A is totally soluble. This result indicates that deoxy-Hb S interacts with deoxy-Hb A and decreases its solubility.  相似文献   

2.
Surface hydrophobicity, stability, solubility, and kinetics of polymerization were studied using hemoglobins with four different amino acids at the beta 6 position: Hb A (Glu beta 6), Hb C (Lys beta 6), Hb Machida (Gln beta 6), and Hb S (Val beta 6). The surface hydrophobicity increased in the order of Hb C, Hb A, Hb Machida, and Hb S, coinciding with the hydrophobicity of the amino acid at the beta 6 position. Solubility of the oxy-form of these hemoglobins decreased in relation to increases in their surface hydrophobicity, suggesting that the solubility is controlled by the strength of hydrophobicity of the amino acid at the beta 6 position. The solubility of the oxy-form of these hemoglobins is always higher than that of the deoxy-form. There is a similar linear relationship between the solubility and surface hydrophobicity among deoxyhemoglobins A, C, and Machida. However, the solubility of deoxy-Hb S deviated significantly from the expected value, indicating that the extremely low solubility of deoxy-Hb S is not directly related to the hydrophobicity of the beta 6 valine. Kinetic studies on the polymerization of deoxy-Hb Machida revealed a distinct delay time prior to polymerization. This confirms our previous hypothesis that beta 6 valine is not responsible for the delay time prior to gelation. The kinetics of the polymerization of 1:1 mixtures of sickle and non-sickle hemoglobins were similar to those of pure Hb S, suggesting that only one of the two beta 6 valines is involved in an intermolecular contact. In mixtures of equal amounts of Hb S and Hb A, Hb C, or Hb Machida, half of the asymmetrical AS, SC, and S-Machida hybrid hemoglobins behaved like Hb S during nucleation, while the other half behaved like the non-sickle hemoglobin.  相似文献   

3.
Sickle hemoglobin (Hb S) was cross-linked by two types of bifunctional imidoesters, dimethyladipimidate (DMA) and dimethyl-3,3'-dithiobispropionimidate (DTBP). These modified hemoglobins were separated into monomer, dimer and polymer fractions by gel filtration. All of these modified hemoglobins showed extremely left-shifted oxygen equilibrium curves with no cooperativity. The stabilities of these hemoglobins were also decreased. The solubilities of these modified hemoglobins in high-phosphate buffers were lower than those of native Hb S. Studies on the kinetics of the aggregation of these modified hemoglobins showed that intracross-linked Hb S with DMA and DTBP (DMA- and DTBP-modified monomeric Hb S) still retained the capability of aggregation with a delay time, while intercross-linked Hb S with DMA and DTBP (DMA- and DTBP-modified oligomeric Hb S) aggregated without a delay time. When the kinetics of aggregation was measured for mixtures of modified and native deoxy-Hb S, DMA-modified monomeric deoxy-Hb S shortened the delay time prior to aggregation of native deoxy-Hb S. The other modified deoxy-Hb S did not affect the delay time, suggesting that these modified oligomeric hemoglobins neither participate in the formation of nuclei nor copolymerize with native deoxy-Hb S.  相似文献   

4.
The role of Asp-beta 73 on the surface hydrophobicity and solubility of hemoglobin was studied using Hb A, Hb S, Hb C Harlem (alpha 2 beta 2Val-6,Asn-73), and Hb Korle Bu (alpha 2 beta 2Asn-73). The surface hydrophobicity of the oxy form of these hemoglobins increased in the order of Hb A, Hb Korle Bu, Hb S, and Hb C Harlem, coinciding with the change in solubility. The same is not true for deoxyhemoglobins. The solubilities of deoxy-Hb S and deoxy-Hb C Harlem were much lower than that expected from their surface hydrophobicity. Although the hydrophobicity of deoxy-Hb C Harlem is greater than that of deoxy-Hb S, the solubility of deoxy-Hb S is only one-third that of deoxy-Hb C Harlem. This deviation must be caused by the substitution of Asn for Asp at the beta 73 position and its inhibitory effect on hydrogen bonding in Hb S polymers. The kinetics of the polymerization of 1:1 mixtures of the deoxy form of S-C Harlem, A-C Harlem, Korle Bu-S, and Korle Bu-C Harlem were studied in comparison with that of deoxy-Hb S and deoxy-Hb C Harlem alone. All of these binary mixtures polymerized with a distinct delay time prior to polymerization. Based on the results of kinetic studies, the probability factors for nucleation of S-C Harlem, A-S, A-C Harlem, S-Korle Bu, and Korle Bu-C Harlem hybrid hemoglobins were calculated as 0.65, 0.5, 0.5, 0.15, and 0.17, respectively, in comparison with that of Hb S (1.0). The probability factor for Hb C Harlem alone was 0.3. These data suggest that the Asp-beta 73 is directly involved in nucleation during Hb S polymerization and that the beta 73 is always trans to the active Val-beta 6 in the formation of nuclei.  相似文献   

5.
Polymerization of half-liganded Hb S was investigated using Ni(II)-Fe(II) hybrid Hb S, in which heme in either alpha or beta s subunits is replaced by Ni (II) protoporphyrin IX. Studies on the polymerization of these hybrid hemoglobins were carried out under aerobic conditions. Both alpha 2 (Ni) beta 2s (Fe-CO) and alpha 2 (Fe-CO) beta 2s (Ni) polymerized with a distinct delay time as do native deoxy-Hb S and Ni(II) Hb S. However, the critical concentration for polymerization of half-liganded Hb S, alpha 2 (Ni) beta 2s (Fe-CO) and alpha 2 (Fe-CO) beta 2s (Ni), was 4- and 8-times higher, respectively, than that of Ni(II)-Hb S. Kinetics of polymerization of both deoxygenated hybrid hemoglobins with CO completely removed were the same, although the critical concentrations for polymerization were intermediate between those for deoxy-Hb S and Ni(II)-Hb S. These results suggest that the small tertiary conformational change associated with the doubly liganded state may be much less favorable to polymerization than the completely unliganded state of Hb S. The conformational change depends on whether alpha or beta chain is liganded. The ease of polymerization and low solubility of sickle hemoglobin is dependent not only on quaternary, but on tertiary structural changes, as well as on the substitution of Val for Glu at the beta 6 position.  相似文献   

6.
We found that hemoglobins A, F and S could be separated on TSK-GEL-SW columns by differences in surface hydrophobicity when eluted with 1.8 M phosphate buffer, pH 7.4. The elution pattern of the oxy- and deoxy-forms of hemoglobins A, S and F from a TSK-GEL-SW-type gel filtration column is useful for measuring surface hydrophobicity. The elution volumes of oxyhemoglobins F, A and S on the TSK-GEL-SW column in 1.8 M potassium phosphate buffer, pH 7.4, related linearly to the log of their solubility; the higher the surface hydrophobicity, the lower the solubility. There was no linear relationship between the solubilities and the elution volumes of these hemoglobins in the deoxy-form; deoxy-Hb S was far from the lines formed by deoxy-Hb A and deoxy-Hb F. These data suggest that the solubility of oxyhemoglobins is related to simple hydrophobic interactions caused by the total surface hydrophobicity, but the extremely low solubility of deoxy-Hb S must be the result of a stereospecific strong hydrophobic interaction between amino acids at the contact regions of deoxy-Hb S molecules.  相似文献   

7.
Deoxygenated sickle cell hemoglobin (Hb S) in 1.8 M phosphate buffer, and carbon monoxide (CO) saturated buffer were rapidly mixed using a stopped-flow apparatus. The binding of the CO to the Hb S polymers and the polymer melting was measured by time resolved optical spectroscopy. Polymer melting was associated with decreased turbidity, and CO binding to deoxy-Hb S was monitored by observation of changes in the absorption profile. The reaction temperature was varied from 20 degrees C to 35 degrees C. Polymer domain size at 20 degrees C was also varied. The data for mixtures involving normal adult hemoglobin (Hb A) fit well to a single exponential process whereas it was necessary to include a second process when fitting data involving Hb S. The overall Hb S-CO reaction rate decreased with increasing temperature from 20 degrees C to 35 degrees C, and increased with decreasing domain size. In comparison, Hb A-CO reaction rates increased uniformly with increasing temperature. Two competing reaction channels in the Hb S-CO reaction are proposed, one involving CO binding directly to the polymer and the other involving CO only binding to Hb molecules in the solution phase. The temperature dependence of the contribution of each pathway is discussed.  相似文献   

8.
Polymerization of a 1:1 mixture of hemoglobin S (Hb S) and the artificial mutant HbAbeta73Leu produces a dramatic morphological change in the polymer domains in 1.0 M phosphate buffer that are a characteristic feature of polymer formation. Instead of feathery domains with quasi 2-fold symmetry that characterize polymerization of Hb S and all previously known mixtures such as Hb A/S and Hb F/S mixtures, these domains are compact structures of quasi-spherical symmetry. Solubility of Hb S/Abeta73Leu mixtures was similar to that of Hb S/F mixtures. Kinetics of polymerization indicated that homogeneous nucleation rates of Hb S/Abeta73Leu mixtures were the same as those of Hb S/F mixtures, while exponential polymer growth (B) of Hb S/Abeta73Leu mixtures were about three times slower than those of Hb S/F mixtures. Differential interference contrast (DIC) image analysis also showed that fibers in the mixture appear to elongate between three and five times more slowly than in equivalent Hb S/F mixtures by direct measurements of exponential growth of mass of polymer in a domain. We propose that these results of Hb S/Abeta73Leu mixtures arise from a non-productive binding of the hybrid species of this mixture to the end of the growing polymer. This "cap" prohibits growth of polymers, but by nature is temporary, so that the net effect is a lowered growth rate of polymers. Such a cap is consistent with known features of the structure of the Hb S polymer. Domains would be more spherulitic because slower growth provides more opportunity for fiber bending to spread domains from their initial 2-fold symmetry. Moreover, since monomer depletion proceeds more slowly in this mixture, more homogeneous nucleation events occur, and the resulting gel has a far more granular character than normally seen in mixtures of non-polymerizing hemoglobins with Hb S. This mixture is likely to be less stiff than polymerized mixtures of other hybrids such as Hb S with HbF, potentially providing a novel approach to therapy.  相似文献   

9.
The abnormal human hemoglobin Malm? (beta97FG4 His leads to Gln) has been studied and its properties are compared with those of normal adult hemoglobin A. The data presented here show that the ring-current shifted proton resonances of both HbCO and HbO2 Malm? are very different from the corresponding forms of Hb A. The hyperfine shifted proton resonances of deoxy-Hb Malm? do not differ drastically from those of deoxy-Hb A. This result, together with the finding that the exchangeable proton resonances of the deoxy form of the two hemoglobins are similar, suggests that unliganded Hb Malm? can assume a deoxy-like quaternary structure both in the absence and presence of organic phosphates We have also compared the properties of Hb Malm? with those of Hb Chesapeake (alpha92FG4 Arg leads to Leu). This allows us to study the properties of two abnormal human hemoglobins with mutations at homologous positions of the alpha and beta chains in the three-dimenstional structure of the hemoglobin molecule. Our present results suggest that the mutaion at betaFG4 has its greatest effect on the teritiary structure of the heme pocket of the liganded forms of the hemoglobin while the mutation at alphaFG4 alters the deoxy structure of the hemoglogin molecule but does not alter the teriary structure of the heme pockets of the liganded form of the hemoglobin molecule. Both hemoglobins undergo a transition from the deoxy (T) to the oxy (R) quaternary structure upon ligation. The abnormally high oxygen affinities and low cooperativities of these two hemoglobins must therefore be due to either the structural differences which we have observed and/or to an altered transition between the T and R structures.  相似文献   

10.
Sickle cell hemoglobin (Hb S) treated with nitrogen mustard (bis(beta-chloroethyl)methylamine hydrochloride) gives two reaction products, one labile and one stable. After dialysis against buffer solution, the remaining stable product is found to inhibit the polymerization of deoxyhemoglobin S. High resolution proton nuclear magnetic resonance has been used to study the structure and function of this stable product and to investigate the nature of the binding sites of nitrogen mustard to the hemoglobin molecule. The NMR results suggest that the nitrogen mustard treatment of Hb S does not alter the heme environment or the subunit interfaces of the hemoglobin molecule. Moreover, the NMR spectra have also shown that the nitrogen mustard reacts with the beta2 histidines of the hemoglobin molecule and have suggested that several other surface amino acid residues of the hemoglobin molecule are also affected by the nitrogen mustard alkylation. These NMR findings are in good agreement with the data obtained from biochemical studies of nitrogen mustard-treated Hb S. The NMR spectra also indicate that nornitrogen mustard (which is also effective in inhibiting sickling) binds with the hemoglobin molecule in a manner identical with nitrogen mustard. Sulfur mustard, on the other hand, produces no observable changes in the aromatic proton resonances, which is consistent with the fact that it does not inhibit the polymerization of deoxy-Hb S.  相似文献   

11.
Sickle cell disease is caused by a mutant form of hemoglobin, hemoglobin S, that polymerizes under hypoxic conditions. The extent and mechanism of polymerization are thus the subject of many studies of the pathophysiology of the disease and potential treatment strategies. To facilitate such studies, a model system using high concentration phosphate buffer (1.5 M-1.8 M) has been developed. To properly interpret results from studies using this model it is important to understand the similarities and differences in hemoglobin S polymerization in the model compared to polymerization under physiological conditions. In this article, we show that hemoglobin S and normal adult hemoglobin, hemoglobin A, aggregate in high concentration phosphate buffer even when the concentration of hemoglobin is below the solubility defined for polymerization. This phenomenon was not observed using 0.05 M phosphate buffer or in another model system we studied that uses dextran to enhance polymerization. We have used static light scattering, dynamic light scattering, and differential interference contrast microscopy to confirm aggregation of deoxygenated and oxygenated hemoglobins below their solubility and have shown that this aggregation is not observable using turbidity measurements, a common technique for assessing polymerization. We have also shown that the aggregation increases with increasing temperature in the range of 15 degrees -37 degrees C and that it increases as the concentration of phosphate increases. These studies contribute to the working knowledge of how to properly apply studies of hemoglobin S polymerization that are conducted using the high phosphate model.  相似文献   

12.
Studies on the aggregation of deoxy-Hb S in concentrated phosphate buffer revealed the formation of three types of polymers, the difference depending on the method employed for polymerization: 1) random or linear polymers without birefringence, 2) helical polymers with birefringence, and 3) crystals. Random or linear polymers were formed when oversaturated deoxy-Hb S was polymerized by the so-called salting out or isothermal method. Helical polymers were formed when oversaturated deoxy-Hb S (120% of the solubility) was polymerized by the temperature jump method. Crystals were formed preferentially by agitation of the sample during the polymerization below 12 degrees C. The solubilities of deoxy-Hb S measured after preparation of these three types of polymers were different, as were the activation energies for the formation of the three polymers. When a mixture of deoxy- and CO-Hb S was crystallized, the crystalline phase did not contain CO-Hb S molecules. To study the relationship among these three types of polymers and red cell sickling, the morphology of erythrocytes was studied after deoxygenation by several different methods. When erythrocytes were prepared by deoxygenation with 2% sodium dithionite at 30 degrees C, a condition similar to that for the isothermal method, red cells did not form the typical sickle shape but rather an irregular shape. In contrast, with the same experiments carried out by using the temperature jump method, typical sickle-shaped cells were formed. These data suggest that the morphological difference may be attributed to the different types of polymers formed inside erythrocytes.  相似文献   

13.
I M Russu  A K Lin  C P Yang  C Ho 《Biochemistry》1986,25(4):808-815
High-resolution proton nuclear magnetic resonance spectroscopy and relaxation techniques have been used to investigate the interactions of sickle cell hemoglobin (Hb S) and human normal adult hemoglobin (Hb A) with p-bromobenzyl alcohol, L-phenylalanine, L-tryptophan, and L-valine. With the exception of valine, all these compounds inhibit the polymerization of deoxy-Hb S [Noguchi, C. T., & Schechter, A. N. (1978) Biochemistry 17, 5455)). Using transferred nuclear Overhauser effects among the proton resonances of the compound of interest and the corresponding longitudinal relaxation rates (T1(-1], we have shown that the binding of each of the compounds investigated to deoxy-Hb S is comparable to that to deoxy-Hb A. Intermolecular transferred nuclear Overhauser effects have been observed between proton resonances of the anti-sickling compounds and specific protons situated in the heme pockets of Hb. On the basis of these results, we suggest that one binding site, common to all compounds with anti-sickling activity, is at or near the heme pockets in the alpha and beta chains of both deoxy-HB S and deoxy-Hb A. The proton T1(-1) values of the histidyl residues situated over the surface of the hemoglobin molecule indicate that a second binding site is located at or near the beta 6 position, containing the mutation in Hb S (beta 6Glu----Val). The binding of the compounds investigated to the latter site induces conformational changes in the amino-terminal domains of the beta chains.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
In an effort to clarify the role of Glu-beta 121 of Hb S molecules in polymerization, we studied the solubility and kinetics of polymerization of various mixtures of deoxyhemoglobins S (Glu-beta 6----Val) and D Los Angeles (Glu-beta 121----Gln). It is known that patients with Hb S-D Los Angeles have a relatively severe clinical course. Mixtures of Hb S and Hb D Los Angeles polymerized after a distinct delay time, the length of which depended on the initial hemoglobin concentration and the fraction of Hb S in the mixture. There was a linear relationship between the logarithmic plot of delay time and initial hemoglobin concentration. The line for a 1:1 mixture of Hb S and Hb D Los Angeles shifted to the right of that for deoxy-Hb S by 0.08. This shift is much smaller than the shift of 0.32 for 1:1 AS mixtures. From these data, the probability factor for nucleation of S-D Los Angeles hybrid hemoglobin was calculated to be 1.16, which is higher than that of Hb S (1.0) and AS hybrid hemoglobin (0.5). The degree of co-polymerization of Hb D Los Angeles in S-D Los Angeles mixtures was similar to that of Hb A in AS mixtures. The critical concentration for the polymerization of Hb D Los Angeles was between that of Hb A and Hb Machida, which has the same amino acid substitution (Glu----Gln) at the beta 6 position. These results suggest that the protein interaction of Hb S molecules during nucleation involves at least two steps. First, the Val-beta 6 of a Hb S molecule interacts hydrophobically with the Phe-beta 85 and the Leu-beta 88 of an adjacent Hb S molecule. In the second step, Glu-beta 121 weakens the interaction with His-beta 116 and Pro-alpha 114. The substitution of Glu-beta 121----Gln may strengthen this second reaction and facilitate nucleation as well as polymerization.  相似文献   

15.
Specific ligation states of hemoglobin are, when crystallized, capable of taking on multiple quaternary structures. The relationship between these structures, captured in crystal lattices, and hemoglobin structure in solution remains uncertain. Wide-angle X-ray solution scattering (WAXS) is a sensitive probe of protein structure in solution that can distinguish among similar structures and has the potential to contribute to these issues. We used WAXS to assess the relationships among the structures of human and bovine hemoglobins in different liganded forms in solution. WAXS data readily distinguished among the various forms of hemoglobins. WAXS patterns confirm some of the relationships among hemoglobin structures that have been defined through crystallography and NMR and extend others. For instance, methemoglobin A in solution is, as expected, nearly indistinguishable from HbCO A. Interestingly, for bovine hemoglobin, the differences between deoxy-Hb, methemoglobin and HbCO are smaller than the corresponding differences in human hemoglobin. WAXS data were also used to assess the spatial extent of structural fluctuations of various hemoglobins in solution. Dynamics has been implicated in allosteric control of hemoglobin, and increased dynamics has been associated with lowered oxygen affinity. Consistent with that notion, WAXS patterns indicate that deoxy-Hb A exhibits substantially larger structural fluctuations than HbCO A. Comparisons between the observed WAXS patterns and those predicted on the basis of atomic coordinate sets suggest that the structures of Hb in different liganded forms exhibit clear differences from known crystal structures.  相似文献   

16.
Asymmetrical cross-linked FS (alpha 2 gamma beta s) hybrid hemoglobin (Hb FS-fumarate) was prepared by reacting mixtures of hemoglobins F and S with double-headed aspirin, bis(3,5-dibromosalicyl) fumarate. When the molar ratio of hemoglobin to the cross-linking agent was 1 to 2 in a 1:1 FS mixture, the relative ratio of the products, cross-linked hemoglobins F (Hb F-fumarate), FS (HB FS-fumarate), and S (Hb S-fumarate), was 1.0:2.6:2.0, in contrast to a 1:2:1 ratio of cross-linked hemoglobins A, AS, and S in a 1:1 AS mixture. These results suggest that the fumaryl group reacts differently with Hb F, Hb FS and Hb S, and that the difference could be attributed to the difference in the structure in the vicinity of the EF6 Lys of non alpha-chains. The oxygen-binding properties of Hb F-fumarate, Hb FS-fumarate, and Hb S-fumarate were similar, except that the n-value of Hb F-fumarate was slightly lower than n-values of Hb S-fumarate and Hb FS-fumarate. Kinetic studies on aggregation showed that the addition of Hb FS-fumarate to unmodified Hb S did not affect the delay time prior to aggregation, but did increase the total turbidity. Electrophoretic and densitometric scanning analysis of the aggregate phase of this mixture showed the fraction of Hb FS-fumarate to be 19%. Hb F-fumarate's effect on the delay time is concentration-dependent; the greater the concentration of Hb F-fumarate, the longer the delay time. The turbidity after aggregation of the mixture of Hb S and Hb F-fumarate was much less than that of Hb S and Hb FS-fumarate. However, the fraction of Hb F-fumarate in the aggregate phase was 19%, which is similar to that of Hb FS-fumarate. These data suggest that Hb F and FS hybrid hemoglobin cannot participate in nuclei formation, but can participate in aggregation after sufficient amounts of nuclei are formed from Hb S, and that increased levels of Hb F do not have an inhibitory effect on the formation of nuclei but on the growth of aggregates.  相似文献   

17.
The kinetics of aggregation and the solubility of deoxy Hb2 CHarlem (α2β2 6 Val, 73 Asn) in concentrated phosphate buffers were studied in comparison with those of deoxy Hb S and deoxy Hb A. Deoxy Hb CHarlem aggregated with a clear exhibition of a delay time. The length of the delay and aggregation times and the degree of the aggregation depended upon the initial hemoglobin concentration.The initial hemoglobin concentration required for the aggregation of deoxy Hb CHarlem was approximately 200% of its solubility, a value much higher than that required for the aggregation of deoxy Hb S (120%). With the same hemoglobin concentration, the delay time for the aggregation of deoxy Hb CHarlem was approximately 100 times longer than that of deoxy Hb S. The logarithmic plotting of the delay time versus hemoglobin concentration in 1.8 m-phosphate buffer (pH 7.4) showed linear lines with a slope (n) of 4.0 for deoxy Hb CHarlem. In contrast to the results for the aggregation of deoxy Hb S, n values for deoxy Hb CHarlem were unchanged with phosphate concentrations varying from 1.2 m to 2.0 m. The solubilities of deoxy Hb S and deoxy Hb CHarlem were increased exponentially by lowering the pH of the medium, with the increase being more conspicuous for Hb CHarlem. The gels (or aggregates) of Hb CHarlem were converted to crystals at a rate much faster than were those of Hb A and Hb S. The kinetics for gelation and crystallization of deoxy Hb CHarlem can be explained by the following scheme, where nuclei G and nuclei C are formed before gelation and crystallization, respectively. Monomenc deoxy Hb
The hemoglobin concentration required for the crystallization of deoxy Hb CHarlem was about ten times lower than that required for deoxy Hb A. The solubility of deoxy Hb CHarlem after aggregation was about twice that of deoxy Hb S, suggesting that the substitution of Asn for Asp at the β73 residue inhibits the formation of nuclei G and accelerates the formation of nuclei C.  相似文献   

18.
Asymmetrical hybrid hemoglobins formed in mixtures of Hb A and Hb S, Hb F and Hb S, Hb S and Hb York(beta 146 His----Pro), and Hb A and Hb York were separated by high performance liquid chromatography on cation and anion exchange columns under anaerobic conditions. The ratio of the hybrid hemoglobin to the total mixture was consistently lower than that theoretically expected and decreased with longer elution times. The hybrid tetramer appears to be unstable even under anaerobic conditions and dissociates into alpha beta dimers. The time course of dissociation of the hybrid hemoglobins was determined by varying the separation programs and thus separating the hybrid hemoglobin at different elution times. The rate of the dissociation of the hybrid hemoglobins studied follows first order kinetics. The lines representing the time course of dissociation of hybrid hemoglobins were extrapolated to time 0 to determine the fraction of the hybrid hemoglobin in the mixture prior to separation. The values obtained for equimolar mixtures of Hb A and Hb S and Hb York and Hb S or Hb A were in agreement with the expected theoretical value (50%). In contrast, the value obtained for hybrid hemoglobin FS was slightly less (about 40%). AY and SY hybrid hemoglobins dissociated into dimers at a considerably faster rate than did AS and FS hybrid hemoglobins, possibly because of the mutation at the beta 146-position in hybrid hemoglobins containing alpha beta Y dimers. This mutation hinders the formation of salt bridges that normally stabilize the "T" quaternary conformation. Since such hybrid hemoglobins have a partial "R" conformation even when deoxygenated, their rate of dissociation to dimers is expected to increase.  相似文献   

19.
Hemoglobins (Hb) Yakima and Kempsey were purified from patients' blood with diethylaminoethyl cellulose column chromatography. The oxygen equilibrium curves of the two hemoglobins and the effects of organic phosphates on the function were investigated. In 0.1 M phosphate buffer, Hill's constants n for Hb Yakima and Hb Kempsey were 1.0 to 1.1 at the pH range for 6.5 to 8.0 and the oxygen affinities of both the mutant hemoglobins were about 15 to 20 times that of Hb A at pH 7.0. The Bohr effect was normal in Hb Yakima and one-fourth normal in Hb Kempsey. In the presence of inositol hexaphosphate, the oxygen affinities to Hb Yakima and Hb Kempsey were greatly decreased, and an interesting result revealed that these hemoglobins showed clear cooperativity in oxygen binding. Hill's constant n in the presence of inositol hexaphosphate was 1.9 for Hb Kempsey and 2.3 for Hb Yakima at pH 7.0. The cooperativities of these mutant hemoglobins were pH-dependent, and Hb Kempsey showed high cooperativity at low pH (n equal 2.1 at pH 6.6) and low cooperativity at high pH (n equal 1.0 at pH 8.0). Hb Yakima showed similar pH dependence in cooperativity. In the presence of inositol hexaphosphate, Hb A showed a pH-dependent cooperativity different from those of Hb Yakima and Hb Kempsey, namely, Hill's n was the highest in alkaline pH (n equal 3.0 at pH 8.0) and decreased at lower pH (n equal 1.5 at pH 6.5). 2,3Diphosphoglycerate bound with the deoxygenated Hb Yakima and Hb Kempsey, however, had no effect on the oxygen binding of these abnormal hemoglobin. The pH-dependent cooperativity of alpha1beta2 contact anomalous hemoglobin and normal hemoglobin was explained by the shifts in the equilibrium between the high and low ligand affinity forms.  相似文献   

20.
The polymerization properties of the fully liganded fluoromet derivative of hemoglobin S (FmetHb S) were investigated by electron microscopy and absorption spectroscopy. Polymerization progress curves, as measured by increasing sample turbidity at 700 nm, exhibit a delay time (t(d)) consistent with the double nucleation mechanism. The pattern of fiber growth, as monitored by electron microscopy, is also indicative of a heterogeneous nucleation process, and dimensions of the fibers were found to be comparable to that of deoxyHb S. The polymerization rate constant (1/t(d)) depends exponentially on Hb S concentration, and the size of the homogeneous and heterogeneous nuclei also depend on FmetHb S concentration. As for deoxyHb S, higher concentrations of protein and phosphate favor fiber formation, while lower temperatures inhibit polymerization. Solubility experiments reveal, however, that eight times more FmetHb S is required for polymerization. The current studies further show that reaction order is independent of phosphate concentration if Hb S activity and not concentration is considered. The allosteric effector, inositol hexaphosphate (IHP), promotes fiber formation, and temperature-dependent reaggregation of FmetHb S suggests that IHP stabilizes pregelation aggregates. These studies show that FmetHb S resembles deoxyHb S in many of its polymerization properties; however, IHP-bound FmetHb S potentially provides a unique avenue for future studies of the early stages of Hb S polymerization and the effect of tertiary and quaternary protein structure on the polymerization process.  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