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.
Transverse water proton relaxation times (T2) have been measured as a function of time after deoxygenation of solutions containing hemoglobin S. The shortened T2 values observed upon deoxygenation of hemoglobin S result from an increase in the correlation time (τc) of the water fraction irrotationally bound to deoxyhemoglobin S as it polymerizes. Therefore, the change in τc as a function of time after deoxygenation can be used to measure the rate of polymer formation. The change in τc observed is reasonably fit by the first-order equation τ = τ0 (1 ? e?kt) + τoxy. At a total hemoglobin concentration of approximately 300 mg/ml, the pseudo-first-order rate constant in a heterozygous AS sample is 25 times slower than in a homozygous S sample, k = 0.019 and 0.47 s?1, respectively. Since the transit time for an erythrocyte in vivo is approximately 15 s, these results suggest that the heterozygous A/S erythrocyte would traverse the circulation and become reoxygenated before extensive polymerization and, therefore, cell sickling could occur. For the homozygous S/S erythrocyte, there is ample time for polymerization and for cell sickling during circulation.  相似文献   

2.
We report the results of thermodynamic and kinetic studies on the gelation of mixtures of sickle cell (S) deoxyhemoglobin with normal human adult (A) and fetal (F) deoxyhemoglobins. The delay time of thermally induced gelation was monitored by the increase in turbidity. At the completion of gelation the solubility was determined by sedimenting the polymers and measuring the supernatant concentration spectrophotometrically. Addition of hemoglobins A or F, at mole fractions from 0 to 0.6, resulted in large increases in both the solubility and the delay time. For a 50:50 mixture of deoxyhemoglobin F with deoxyhemoglobin S, the solubility increased by a factor of 1.8 and the delay time by a factor of 107 relative to pure deoxyhemoglobin S at the same total concentration, while for a 50:50 mixture of deoxyhemoglobins A and S the solubility increased by a factor of 1.4 and the delay time by a factor of 104. The relative delay times were independent of both temperature and total hemoglobin concentration. The data have been analyzed according to theoretical models which treat the effects of temperature, concentration, non-ideality and solution composition on the thermodynamics and kinetics of gelation. The increased solubility in mixtures with deoxyhemoglobin F is fully explained by a model in which only deoxyhemoglobin S molecules polymerize. The effect of fetal hemoglobin (α2γ2) and hybrid α2γβS molecules is to increase the solution non-ideality through the contribution of their excluded volume. The smaller increase in the solubility observed in comparable mixtures with deoxyhemoglobin A requires that the hybrid α2βAβS molecules copolymerize with the deoxyhemoglobin S. The kinetic results for the mixtures can be quantitatively accounted for using a nucleation model in which the equilibrium properties of the polymer are used to describe the critical nucleus. The very large increases in delay time observed for the SF mixtures can be explained by assuming that only α2β2S molecules participate in the formation of a nucleus containing about 25 monomers. As in the thermodynamic analysis, the smaller effect of adding deoxyhemoglobin A can be attributed to the contribution of the hybrid molecules in forming the critical nucleus. Thus the difference between the polymerization properties of mixtures of deoxyhemoglobin S with deoxyhemoglobins A and F can be attributed solely to the copolymerization of the α2βAβS hybrid molecule and the absence of any significant copolymerization of the α2γβS hybrid.  相似文献   

3.
K Adachi  T Asakura 《Blood cells》1982,8(2):213-224
Diluted solutions of deoxyhemoglobin S in concentrated phosphate buffer form aggregates or gels with a clear exhibition of a delay time. The aggregates can be liquified by cooling, bubbling with O2 or CO gas, or the dilution of phosphate buffer with water. These properties can be used as a simple method for studying the mechanism of polymerization and depolymerization of hemoglobins. The advantages of this method are: 1) The amount of hemoglobin sample required is only 1% to 5% of that required for the gelation of deoxy-Hb S in low phosphate buffer. 2) The kinetics can be measured turbidimetrically using an ordinary spectrophotometer. 3) The solubility of hemoglobin can be directly determined by taking the absorption spectrum of the supernatant solution after polymerization. 4) The polymer phase can be easily separated from the solution so that the amount and composition of the polymers can be analyzed. 5) The volume of the polymer phase is so small that excluded volume effect can be neglected. 6) The method can be applied to the study of polymerization of non-sickle hemoglobins and that of mixtures of sickle and non-sickle hemoglobins. The major question is whether the polymerization of hemoglobin in concentrated phosphate buffer is the same as that of deoxy-Hb S in low phosphate buffer. To answer this question, we studied the polymerization of Hb S, Hb A, Hb C Harlem, and Hb C in phosphate buffers of different molarities. We also studied the mechanism of the conversion of gels of these hemoglobins into crystals.  相似文献   

4.
The solubility equilibrium between monomer and polymer which has been shown to exist in deoxyhemoglobin S solutions is examined in solutions partially saturated with carbon monoxide. The total solubility is found to increase monotonically with increasing fractional saturation. At low fractional saturations the increase is nearly linear, amounting roughly to an increase of 0.01 g cm?3 in solubility for each 10% increase in fractional saturation. Linear dichroism measurements on the spontaneously aligned polymer phase are used to examine the composition of the polymer as a function of the fractional saturation of the corresponding solution phase. The dichroism experiments show that the polymer phase contains less than 5% of CO-liganded hemes even at supernatant fractional saturations in excess of 70%. The polymer selects against totally liganded hemoglobin molecules by a minimum factor of 65 and against singly liganded molecules by a factor of at least 2.5. Consequently, polymerized hemoglobin S has a ligand affinity which is significantly lower than that of monomeric hemoglobin S in the deoxy quaternary structure.The kinetics of the polymerization reaction in the presence of CO are similar to those observed in pure deoxyhemoglobin S solutions. The polymerization is preceded by a pronounced delay, the duration of which, td, is proportional roughly to the 30th power of the solubility. At low fractional saturations, this amounts to a tenfold increase in td for each 10% increase in the fractional saturation.These results show that the polymerization reaction is nearly specific for deoxyhemoglobin. Models for the dependence of the solubility and the polymer saturation on ligand partial pressure demonstrate the importance of solution phase non-ideality in determining the solubility of mixtures. The results require selection against partially liganded species which is significantly greater than is predicted by the two-state allosteric model. The data are compatible with either sequential or allosteric models in which the major polymerized component is the unliganded hemoglobin molecule.  相似文献   

5.
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   

6.
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   

7.
The kinetics of gelation of deoxyhemoglobin S were investigated as a function of temperature, concentration of hemoglobin, and solvent composition. Measurements were made by continuously monitoring the changes in viscosity with time, after polymerization had been induced by rapidly raising the temperature. A specially constructed low-shear viscometer was used. The solution density was also measured continuously to determine whether a volume change accompanied aggregation.The results confirm earlier work in showing that the time-dependence of the viscosity is composed of a variable latent period (several minutes to tens of hours) during which there is only a slight and very gradual increase in viscosity, followed by a stage in which the viscosity rises very sharply within a very short time. The length of the initial latent period is highly dependent upon the HbS3 concentration (33rd ± 6 power) and temperature. If the duration is interpreted as the inverse of a reaction rate, the activation energy is 96 ± 10 kcal/mol for solutions containing inosital hexaphosphate. Unlike measurements reported by others, no dependence of the latent period on shear rate was observed at the low shear rate employed. When IHP is omitted from the hemoglobin solutions, qualitatively similar results are obtained; however, the latent period depends on the 26th ± 6 power of the deoxyhemoglobin S concentration and yields an average activation energy of 125 ± 10 kcal/mol. The length of the latent period is increased 40-fold. Tris is known to prevent gelation but the inhibition can be partly reversed by adding IHP. When this is done, highly concentration-dependent latent periods are again observed. The results may be interpreted in terms of nucleation kinetic theories: a critical nucleus composed of approximately 30 hemoglobin molecules is required for gelation; and the energy barrier (which is larger in the absence of IHP) to the formation of this critical aggregate is approximately 100 kcal/mol.Gelation is not accompanied by a detectable volume change (limits 5 × 10?5 g/ml). This indicates that the volume change of the reaction must be less than + 60 cm3/mol when the aggregates represent one half of the HbS available for polymerization.  相似文献   

8.
Using a combination of laser photolysis and temperature-jump techniques, the kinetics of hemoglobin S polymerization have been studied over a wide range of delay times (10(-3) to 10(5)s), concentrations (0.2 to 0.4 g/cm3) and temperatures (5 to 50 degrees C). A slow temperature-jump technique was used to induce polymerization in samples with delay times between 10(2) seconds and 10(5) seconds by heating a solution of completely deoxygenated hemoglobin S. For samples with shorter delay times, polymerization was induced by photodissociating the carbon monoxide complex in small volumes (10(-9) cm3) using a microspectrophotometer equipped with a cw argon ion laser. The photolysis technique is described in some detail because of its importance in studying hemoglobin S polymerization at physiological concentrations and temperatures. In order, to establish conditions for complete photodissociation with minimal laser heating, a series of control experiments on normal human hemoglobin was performed and theoretically modeled. The concentration dependence of the tenth time is found to decrease with increasing hemoglobin S concentration. In the range 0.2 to 0.3 g/cm3, the tenth time varies as the 36th power of the hemoglobin S concentration, while in the range 0.3 to 0.4 g/cm3 it decreases to 16th power. As the tenth times become shorter, the progress curves broaden, with the onset of polymerization becoming less abrupt. For tenth times greater than about 30 seconds, measurements with the laser photolysis technique on small volumes yield highly irreproducible tenth times, but superimposable progress curves, indicating stochastic behavior. The initial part of the progress curves from both temperature-jump and laser photolysis experiments is well fit with an equation for the concentration of polymerized monomer, delta (t) = A[cosh (Bt) -1], which results from integration of the linearized rate equations for the double nucleation mechanism described in the accompanying paper (Ferrone et al., 1985). The dependence of the parameters A and B on temperature and concentration is obtained from fitting over 300 progress curves. The rate B has a large concentration dependence, varying at 25 degrees C from about 10(-4) S-1 at 0.2 g/cm3 to about 100 s-1 at 0.4 g/cm3.  相似文献   

9.
From a consideration of the primary sequence of hemoglobin A2 and the reported 5 å molecular contacts between deoxyhemoglobin S molecules in a crystal, it is predicted that hemoglobin A2 might act as an inhibitor of the polymerization of deoxyhemoglobin S in a manner similar to hemoglobin. F. This has been tested experimentally by measuring the rate of change of the transverse water proton relaxation times (T2) in equimolar mixtures of hemoglobin S and one of the non-gelling hemoglobins A, F or A2. Hemoglobins A2 and F have far more pronounced inhibitory effects on the rate of polymerization than does hemoglobin A. These molecules contain several amino acid differences from hemoglobin A beta chains which are located in the 5 Å molecular crystal contacts and these altered crystal contacts result in a much stronger inhibition of the rate of polymerization. Since hemoglobin A2 is a normal hemoglobin found in small amounts in all adult red cells, increased delta chain synthesis may have potential importance in therapy for sickle cell disease.  相似文献   

10.
Two assays of equilibrium properties in the gelation of deoxyhemoglobin S were carried out by analytical ultracentrifugation on the same sample: Csat, the monomer concentration in equilibrium with the fully formed gel, was obtained as the supernatant concentration after sedimentation of a preformed gel. The presence of a plateau region during sedimentation of the supernatant and the rate of sedimentation of the boundary from which Csat was measured indicate that centrifugation did not alter the pre-existing equilibrium and that the supernatant consisted of monomers. The centrifugation was then continued to equilibrium to obtain a distribution showing a sharp increase in molecular weight at Cagg, the monomer concentration at which a small amount of polymerization to large aggregates begins.The primary result is that Csat > Cagg under all conditions. The different values of the two parameters indicate that they reflect two separate transitions and that the overall monomer to gel process has a limited co-operativity. Within the limits of the method Csat is independent of total hemoglobin concentration. The two transitions divide the overall range of total hemoglobin concentration into an essentially monomeric region at concentrations below Cagg, a region in which isotropically oriented polymers exist, occurring when monomer concentration lies between Cagg and Csat, and a two-phase region of conjugate isotropic and anisotropic phases when monomer concentration equals Csat. These regions correspond to zones in the ultracentrifuge equilibrium distribution. In this scheme Cagg depends only on the interaction energy of polymerization. Csat depends on entropic factors which induce tactoid formation as well. Csat, while a monomer concentration, reflects a saturation not of monomers in relation to a polymeric phase, but of polymers in the isotropic phase in relation to the anisotropic or tactoidal polymerized phase. As such, Csat represents a supersaturated state of isolated monomers.The ratio CsatCagg = 1.23 in stripped hemoglobin3 and equilibrium distributions in the zone of isotropically oriented polymers were both used to obtain an order of magnitude estimate of polymer size, found to be much smaller than that of hemoglobin S fibers. This further confirms that gelation does not consist of a single transition and phase change with near infinite co-operativity of polymerization.Csat as well as Cagg are lowered by 2,3,diphosphoglycerate and inositol hexa-phosphate. Decreasing pH near 7 also favors gelation; in stripped hemoglobin a pH optimum for gelation occurs near pH 6.8. The apparent van't Hoff ΔH for stripped hemoglobin is about 3 kcal/mol for Csat and 2 kcal/mol for Cagg.  相似文献   

11.
The temperature and cell volume dependence of the NMR water proton linewidth, 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 dexygenation 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 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°C to 500 Hz at ?36°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°C for measurements at 44.4 MHz and ?50°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 irratationally bound water, is altered during the sickling process.  相似文献   

12.
The final step in the erythrocyte methemoglobin reduction pathway, the transfer of an electron from cytochrome b5, to methemoglobin, has been studied using magnetic circular dichroism spectroscopy. Spectral analysis allowed us to determine accurately the concentration of each redox species in mixtures of the two heme-proteins and to follow simultaneously the kinetics of the appearance or disappearance of each of these species during reduction reactions. Our analysis detected a substantial increase in the high-spin hydroxymethemoglobin species in the partially reduced bovine hemoglobin tetramer. This species was sensitive to the degree of reduction and pH, and was spectrally similar to fluoride methemoglobin. At pH 7.8. 100% of the hydroxide component of methemoglobin was in the high-spin form when two or more subunits were in the ferrous form. Kinetic analysis of bovine methemoglobin reduction yielded values for the apparent first-order rates for the tetrameric species possessing four, three, two, and one ferric subunit. Further analysis showed that the reduction kinetics can also be described by an equilibrium state, pure competitive inhibition model for enzyme catalysis in which ferrous and ferric subunits of hemoglobin compete for cytochrome b5 This analysis generated a KD that depends on ionic strength and hemoglobin tetramer conformation, a Vmax that was independent of these factors, and an inhibition constant that was equal to Kd. This model is consistent with the hypothesis that the reduction of methemoglobin can be separated into two steps, the ionic interaction between cytochrome b5 and hemoglobin and the electron transfer.  相似文献   

13.
A double nucleation mechanism for the polymerization of sickle hemoglobin is described. The mechanism accounts for all of the major kinetic observations: the appearance of a delay, the high concentration dependence of the delay time, and the stochastic behavior of slowly polymerizing samples in small volumes. The mechanism postulates that there are two pathways for polymer formation: polymerization is initiated by homogeneous nucleation in the solution phase, followed by nucleation of additional polymers on the surface of existing ones. This second pathway is called heterogeneous nucleation. Since the surface of polymers is continuously increasing with time, heterogeneous nucleation provides a mechanism for the extreme autocatalysis that is manifested as an apparent delay in the kinetic progress curves. In this mechanism, each spherulitic domain of polymers is considered to be initiated by a single homogeneous nucleation event. The mechanism explains the irreproducibility of the delay time for single domain formation as arising from stochastic fluctuations in the time at which the homogeneous nucleus for the first polymer is formed. Integration of the linearized rate equations that describe this model results in a simple kinetic form: A[cosh(Bt)-1] (Bishop & Ferrone, 1984). In the accompanying paper (Ferrone et al., 1985) it was shown that the initial 10 to 15% of progress curves, with delay times varying from a few milliseconds to over 10(5) seconds, is well fit by this equation. In this paper, we present an approximate statistical thermodynamic treatment of the equilibrium nucleation processes that shows how the nucleus sizes and nucleation equilibrium constants depend on monomer concentration. The equilibrium model results in expressions for B and B2A as a function of monomer concentration in terms of five adjustable parameters: the bimolecular addition rate of a monomer to the growing aggregate, the fraction of polymerized monomers that serve as heterogeneous nucleation sites, the free energy of intermolecular bonding within the polymer, and two parameters that describe the free energy change as a function of size for the bonding of the heterogeneous nucleus to a polymer surface. This model provides an excellent fit to the data for B and B2A as a function of concentration using physically reasonable parameters. The model also correctly predicts the time regime in which stochastic behavior is observed for polymerization in small volumes.  相似文献   

14.
Gelation experiments with artificially formed half-liganded hybrid tetramers of hemoglobin S demonstrate that when either the α chains or the βs chains are fixed in the cyanmet (CNmet) liganded state, gelation occurs upon deoxygenation of the ferrous chains. The minimum concentration of hemoglobin required for gelation is equivalent for both hybrids (α2cnmetβ2s and α2β2scnmet), is considerably higher than the concentration required to gel deoxy-Hb S (α2β2s), and can be restored to the lower minimum gelling point of α2β2s by reduction of the CNmet chains with dithionite. These results suggest that the most important conformational determinant of the deoxy state for polymerization of Hb S is the quaternary deoxy structure rather than the tertiary structural effect of the ligand state of the α or the βs chains, and are furthermore consistent with the notion that asymmetric deoxy-CNmet hybrid tetramers assume a conformation which resembles, but is not identical to that of deoxyhemoglobin.The results of gelation experiments with mixtures of hemoglobins S and A in which selected chains of one or both hemoglobins are in the CNmet form support the concept that certain non-S hemoglobins may participate in the sickling process by forming hybrid tetramers with Hb S (such as α2βaβs). The conformational requirement for participation of these hybrids in polymers also appears to be a quaternary deoxy-like structure.  相似文献   

15.
Hemoglobin (Hb) S containing Leu, Ala, Thr, or Trp substitutions at beta 85 were made and expressed in yeast in an effort to evaluate the role of Phe-beta 85 in the acceptor pocket during polymerization of deoxy Hb S. The four Hb S variants have the same electrophoretic mobility as Hb S, and these beta 85 substitutions do not significantly affect heme-globin interactions and tetramer helix content. Hb S containing Trp-beta 85 had decreased oxygen affinity, whereas those with Leu-, Ala-, and Thr-beta 85 had increased oxygen affinity. All four supersaturated beta 85 variants polymerized with a delay time as does deoxy Hb S. This is in contrast to deoxy Hb S containing Phe-beta 88, Ala-beta 88, Glu-beta 88, or Glu-beta 85, which polymerized with no clear delay time (Adachi K, Konitzer P, Paulraj CG, Surrey S, 1994, J Biol Chem 269:17477-17480; Adachi K, Reddy LR, Surrey S, 1994, J Biol Chem 269:31563-31566). Leu substitution at beta 85 accelerated deoxy Hb S polymerization, whereas Ala, Thr, or Trp substitution inhibited polymerization. The length of the delay time and total polymer formed for these beta 85 Hb S variants depended on hemoglobin concentration in the same fashion as for deoxy Hb S: the higher the concentration, the shorter the delay time and the more polymer formed. Critical concentrations required for polymerization of deoxy Hb SF veta 85L, Hb SF beta 85A, Hb SF beta 85T, and Hb SF beta 85W are 0.65-, 2.2-, 2.5- and 3-fold higher, respectively, than Hb S.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The reactions of human hemoglobin with p-nitro- and p-chlorobenzenediazonium tetrafluoroborates in the presence and absence of molecular oxygen have been investigated in kinetic detail. The oxidation of iron(II) occurs with first order rate dependence on both the hemoglobin and diazonium salt concentrations, but inverse first order dependence on the concentration of molecular oxygen characterizes reactions performed in the presence of O2. In the absence of O2, nitrobenzene is the only product observed from hemoglobin oxidation by p-NO2C6H4N2+BF4?, and a 1:1 stoichiometry exists between nitrobenzene produced and Fe(II) oxidized. In the presence of O2, p-nitrophenol is the dominant product, but product yield is dependent on the ratio of reactants. Electron transfer to the diazonium salt rather than its corresponding diazohydroxide or diazoate is inferred from the relative absence of pH dependence on the rate of oxidation. The composite results are consistent with a mechanism for hemoglobin oxidation that requires molecular oxygen dissociation from oxyhemoglobin prior to oxidation by the diazonium salt. Implications of this investigation for the mechanism of arylhydrazine reactions with hemoglobin are discussed.  相似文献   

17.
18.
The kinetics of proton release on ligation of menhaden hemoglobin was studied by flash photolysis over a range of pH. In contrast to all previous kinetic work with human hemoglobin, a nonlinear relationship between proton release and CO binding was found. Proton uptake was also observed in the course of O2 replacement by CO at low pH. It follows that at least part of the proton release is associated with quaternary rather than tertiary conformational changes i.e. this result is consistent with a two-state model in which L is a function of pH.  相似文献   

19.
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   

20.
Using the stochastic theory of chemical reactions and the theory of first passage times, a simple analytic expression is derived for the distribution of delay times that has been observed in studies of the polymerization kinetics of sickle hemoglobin under conditions where the polymerization progress curves exhibit stochastic variation. The rate of homogeneous nucleation can be readily extracted from such experiments using this expression. This work constitutes a significant addition to the rather limited number of examples where contact can be successfully made between the stochastic theory of chemical kinetics and experiment.  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