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.
Bis(3,5-dibromosalicyl) fumarate was used to crosslink oxyhemoglobin between Lys 82 beta 1 and Lys 82 beta 2 (Walder, J. A., et al. (1979) Biochemistry 18, 4265) and deoxyhemoglobin between Lys 99 alpha 1 and Lys 99 alpha 2 (Chatterjee R.Y., et al. (1986) J. Biol. Chem. 261, 9929). Thermal denaturations demonstrated that alpha crosslinked hemoglobin (alpha 99XLHb A) has the same stability as the beta crosslinked one (beta 82XLHb A). Both alpha and beta crosslinked methemoglobins have a denaturation temperature in 0.9 M guanidine of 57 degrees C compared to 41 degrees C of Hb A. The second product from the T-state crosslinking reaction was found to be crosslinked between the beta chains by chain separation and amino acid analysis. The possible positions for this crosslink are limited to the bisphosphoglycerate binding site in the three-dimensional structure. Its stability is comparable to that of the alpha 99XLHb A or beta 82XLHb A. These modified hemoglobins are potential blood substitutes.  相似文献   

2.
The effect of 2,3-diphosphoglycerate on the solubility of deoxyhemoglobin S   总被引:1,自引:0,他引:1  
Although highly charged polyanions, such as inositol hexaphosphate, have been clearly shown to decrease the solubility of deoxyhemoglobin S, the effect of 2,3-diphosphoglycerate (DPG), the endogenous allosteric effector within the red cell, has been more controversial. In this work we have compared the effect of DPG on the solubility of native deoxyhemoglobin S and a derivative in which the DPG binding site is blocked by cross-linking the two beta 82 lysine residues. At pH 6.6 and 30 degrees C the solubility of deoxyhemoglobin S was found to be decreased by 15% (i.e., from 18.8 to 16.0 g/dl) in the presence of saturating concentrations of DPG. Under the same conditions DPG had no effect on the solubility of the cross-linked derivative. This result establishes unequivocally that the binding of DPG within the beta cleft directly facilitates the polymerization of deoxyhemoglobin S. Under physiological conditions, the solubility of deoxyhemoglobin S was found to be decreased by 6% in the presence of an equimolar concentration of DPG. A solubility decrease of this magnitude is sufficient to enhance the tendency of SS cells to sickle and may exacerbate the clinical symptoms of sickle cell disease.  相似文献   

3.
Hemoglobin can be cross-linked and converted to a bioconjugate in one step by a "molecular necklace", a reagent that contains two reacting sites and a pendant ligand. The compound to be conjugated is activated as an electrophile. The activated material is then combined with a reagent (3-aminoisophthalic acid) that contains a nucleophilic (amino) site and two latent (carboxyl) sites. The latent sites of the product are activated as 3,5-dibromosalicylates to produce the cross-linker. Illustrative examples of cross-linking are presented with pendant biotin [bis(3,5-dibromosalicyl) N-biotinyl-5-aminoisophthalate] and pendant N-trifluoroacetyl-L-isoleucylglycine [bis(3,5-dibromosalicyl) N-(N-trifluoroacetyl-L-isoleucylglycyl)-5-aminoisophthalate) ]. The resulting modified hemoglobins contain two principal types of cross-link: (beta-Lys-82-beta'-Lys-82) and (alpha-Lys-99-alpha'-Lys-99). The functional properties of the modified hemoglobin containing biotin in a (beta-Lys-82-beta'-Lys-82) cross-link are (pH 7.4, 55 microM heme, 25 degrees C, 0.1 M chloride, and 50 mM Bis-Tris) P(50) = 4.9 Torr, n(50) = 3.0, values which are approximately the same as for native hemoglobin. The results of affinity chromatography of the biotinylated cross-linked hemoglobin using a column of immobilized avidin indicate that the pendant biotin is much less accessible than free biotin. We suggest that the results are consistent with the pendant species being strongly attracted into the hemoglobin environment.  相似文献   

4.
Two double-headed aspirins, bis(3,5-dibromosalicyl) succinate and bis(3,5-dibromosalicyl) fumarate, have been found to be potent acylating agents of intracellular hemoglobin (A or S) in vitro. Furthermore, each of these reagents cross-links beta chains of hemoglobin, probably at the beta cleft. The modified hemoglobins show increased oxygen affinities and reduced gelation or sickling tendencies.  相似文献   

5.
Bis(3,5-dibromosalicyl) fumarate and a number of related bifunctional reagents react preferentially with oxyhemoglobin to cross-link the beta chains within the 2,3-diphosphoglycerate-binding site. In this report we describe a new derivative cross-linked between the alpha chains which is formed specifically in the reaction with deoxyhemoglobin. X-ray crystallographic studies show that the cross-link lies between Lys-99 alpha 1 and Lys-99 alpha 2, spanning the central cavity of the tetramer. Lys-99 alpha 1 and Lys-99 alpha 2 are located within a cluster of charged residues very near the middle of the hemoglobin molecule. In oxyhemoglobin, this site is completely inaccessible to the cross-linking agent. Competition experiments with inositol hexaphosphate indicate that the compound enters the central cavity in deoxyhemoglobin through the cleft between the alpha chains. Despite the presence of the cross-link between the alpha chains, the modified hemoglobin remains highly cooperative. The Hill coefficient for HbXL99 alpha is 2.6. The oxygen affinity of the cross-linked derivative is decreased by approximately 2-fold; at pH 7.0 in the presence of 0.1 M NaCl the P50 is 13.9 mm Hg compared to 6.6 mm Hg for HbA. This difference appears to be due to relatively small changes in both KR, the association constant for binding of oxygen to the R state, and the allosteric constant L. Surprisingly, the isoelectric point of oxyHbXL99 alpha is almost identical to that of oxyHbA, whereas in the deoxy form the isoelectric point of the cross-linked derivative is decreased relative to native hemoglobin as expected due to the loss of the two positive charges of the modified amino groups. In agreement with these findings, the alkaline Bohr effect of HbXL99 alpha is decreased by more than 50%. Earlier studies argue strongly against the possibility that Lys-99 alpha is directly responsible for this large fraction of the Bohr effect in HbA. Analysis of the structure suggests that in the cross-linked derivative Glu-101 beta, which is in close proximity to Lys-99 alpha in oxyhemoglobin, becomes an acid Bohr group.  相似文献   

6.
The binding of carbon dioxide to human hemoglobin cross-linked between Lys alpha 99 residues with bis(3,5-di-bromosalicyl) fumarate was measured using manometric techniques. The binding of CO2 to unmodified hemoglobin can be described by two classes of sites with high and low affinities corresponding to the amino-terminal valines of the beta and alpha chains, respectively (Perrella, M., Kilmartin, J. V., Fogg, J., and Rossi-Bernardi, L. (1975b) Nature 256, 759-761. The cross-linked hemoglobin bound less CO2 than native hemoglobin at all CO2 concentrations in deoxygenated and liganded conformations, and the ligand-linked effect was reduced. Fitting the data to models of CO2 binding suggests that only half of the expected saturation with CO2 is possible. The remaining binding is described by a single affinity constant that for cross-linked deoxyhemoglobin is about two-thirds of the high affinity constant for deoxyhemoglobin A and that for cross-linked cyanomethemoglobin is equal to the high affinity constant for unmodified cyanomethemoglobin A or carbonmonoxyhemoglobin A. The low affinity binding constant for cross-linked hemoglobin in both the deoxygenated and liganded conformations is close to zero, which is significantly less than the affinity constants for either subunit binding site in unmodified hemoglobin. Comparing the low affinity sites in this modified hemoglobin to native hemoglobin suggests that cross-linking hemoglobin between Lys alpha 99 residues prevents CO2 binding at the alpha-subunit NH2 termini.  相似文献   

7.
Bis(3,5-dibromosalicyl) fumarate was used to crosslink hemoglobin both in the oxy and deoxy states. This double headed diaspirin was known to crosslink oxy Hb A selectively between Lys 82 beta 1 and Lys 82 beta 2 (Walder, J. A., et al. (1979) Biochemistry 18, 4265) and deoxy Hb A between Lys 99 alpha 1 and Lys 99 alpha 2 (Chatterjee R. Y., et al. (1986) J. Biol. Chem. 261, 9929). The autoxidation at 37 degrees C of oxy alpha 99 crosslinked hemoglobin was found to be 1.8 times as fast as that of Hb A while that of the oxy beta 82 crosslinked hemoglobin was only 1.2 times as fast. After 5 hours the formation of methemoglobin in the alpha crosslinked Hb A is 21.3% compared to 10.8% in beta crosslinked Hb A and 6.4% in Hb A. These results may effect the proposed use of alpha 99 crosslinked hemoglobin as a blood substitute by demonstrating the need for protection from autoxidation during storage.  相似文献   

8.
Human hemoglobin was reacted with the bifunctional reagent bis(3,5-dibromosalicyl) fumarate to yield a derivative (Hb alpha alpha) crosslinked between the two alpha-chains; when the reaction was carried out with HbA already crosslinked between the two beta-chains by 2-nor-2-formylpyridoxal 5'-phosphate, a doubly crosslinked derivative (Hb alpha alpha beta beta) was obtained. We have observed that both modified hemoglobins are extremely stable up to temperatures of at least 85 degrees C. The carbon monoxide binding kinetics of both crosslinked hemoglobins, studied at temperatures between 15 and 85 degrees C, by means of stopped flow and flash photolysis techniques, show that the ligand-linked allosteric transition is maintained even at the highest temperatures. These results are also relevant to the mechanism of thermal unfolding of human hemoglobin, since they show that dissociation into alpha beta dimers (and exposure of the relatively hydrophobic dimer-dimer interfaces) is an obligatory step in the irreversible denaturation of deoxy and carbon monoxy hemoglobin.  相似文献   

9.
Oxygen equilibrium curves of human hemoglobin Ao (HbAo) and human hemoglobin cross-linked between the alpha chains (alpha alpha Hb) by bis(3,5-dibromosalicyl) fumarate were measured as a function of pH and chloride or organic phosphate concentration. Compared to HbAo, the oxygen affinity of alpha alpha Hb was lower, cooperativity was maintained, although slightly reduced, and all heterotropic effects were diminished. The major effect of alpha alpha-cross-linking appears to be a reduction of the oxygen affinity of R-state hemoglobin under all conditions. However, while the oxygen affinity of T-state alpha alpha Hb was slightly reduced at physiologic chloride concentration and in the absence of organic phosphates, KT was the same for both hemoglobins in the presence of 2,3-diphosphoglycerate (or high salt) and higher for alpha alpha Hb in the presence of inositol hexaphosphate. The reduced O2 affinity arises from smaller binding constants for both T- and R-state alpha alpha Hb rather than through stabilization of the low affinity conformation. All four Adair constants could be determined for alpha alpha Hb under most conditions, but a3 could not be resolved for HbAo without constraining a4, suggesting that the cross-link stabilizes triply ligated intermediates of hemoglobin.  相似文献   

10.
The double-headed aspirin, bis(3,5-dibromosalicyl) fumarate, has been used to crosslink hemoglobin A between Lys 82 beta 1 and Lys 82 beta 2 (J. A. Walder et al. (1979) Biochemistry 18,4265). Denaturation experiments were used to compare the stability of this crosslinked protein to that of hemoglobin A. Thermal denaturations, done in 0.01 M 4-morpholine-propanesulfonic acid, pH 7, containing 0.9 M guanidine to prevent precipitation at high temperatures, were monitored by changes in absorbance between 190 and 650 nm using a diode array spectrophotometer. The sample was heated from 25 to 70 degrees C at 0.3 degrees C/min. The data were analyzed by using both a two-state model and a novel first derivative method. As expected, methemoglobin A had a single, broad transition with a midpoint of 40.7 degrees C. The crosslinked methemoglobin showed a transition at 57.1 degrees C. Two minor transitions, one of which was apparently due to residual unmodified hemoglobin, were also observed in the crosslinked sample. Thus, a single crosslink between only two of the four subunits can lead to a significantly more stable molecule. These results can be explained by Le Chatelier's principle, since crosslinking prevents dissociation of the beta-subunits and, thereby, holds the entire tetramer together.  相似文献   

11.
Diaspirin crosslinked hemoglobin (DCLHb) was analyzed by mass spectrometric-based techniques to identify the protein modifications effected by the crosslinking reaction with bis(3,5-dibromosalicyl) fumarate. DCLHb consists of two principal components. These components were isolated by size-exclusion chromatography and identified by measurement of their molecular weight using electrospray mass spectrometry and subsequent peptide mass mapping and mass spectrometric sequence analysis of their individual digests. Three major RP-HPLC fractions were observed from the major hemoglobin in DCLHb. Their MWs matched the MW of heme, intact hemoglobin beta-chain, and two hemoglobin alpha-chains crosslinked by a fumarate moiety, respectively. The minor HPLC peaks of DCLHb were also separated, and characterized by mass spectrometric methods. These minor components revealed additional details of the structural nature of covalent modification of DCLHb.  相似文献   

12.
L A Dick  G Heibel  E G Moore  T G Spiro 《Biochemistry》1999,38(20):6406-6410
UV resonance Raman difference spectra between ligated and deoxyhemoglobin contain tryptophan and tyrosine signals which arise from quaternary H-bonds in the T state, which are broken in the R state. These H-bonds are unaffected by bis(3,5-dibromosalicyl) fumarate cross-linking at the Lys alpha 99 residues, which prevents dissociation of Hb tetramers to dimers. However, when the pH is lowered from 9.0, or when NaCl is added, intensity is diminished for the tyrosine Y8 and tryptophan W3 bands of cross-linked deoxyHb, but not of native deoxyHb. This effect is attributed to weakening of tertiary H-bonds involving Tyr alpha 140 and Trp alpha 14, when the T state salt bridge between Val alpha 1 and Arg alpha 141 is formed via protonation of the terminal amino group and anion binding. The Tyr alpha 140-Val alpha 93 H-bond connects the Arg alpha 141-bearing H helix with the Lys alpha 99-bearing G helix. Weakening of the H-bond reflects a tension between the fumarate linker and the salt-bridge. This tension inhibits protonation of the Val alpha 1 amino terminus, thus accounting for the diminution of both proton [Bohr effect] and CO2 binding in the T state as a result of cross-linking.  相似文献   

13.
Hemoglobins A and S were crosslinked between Lys 82 beta 1 and Lys 82 beta 2 using bis (3,5-dibromosalicyl) fumarate (J. A. Walder et al. (1979) Biochemistry 18, 4265). Thermal denaturation experiments were used to compare the stabilities of the met, cyanomet, and carbonmonoxy forms of these crosslinked hemoglobins to the corresponding uncrosslinked proteins. Uncrosslinked carbonmonoxy- and cyanomethemoglobins had transition temperatures about 11 degrees C higher than the corresponding met samples. The increase in denaturation temperature (Tm) due to crosslinking was 15 degrees C for the methemoglobins, 10 degrees C for the cyanomethemoglobins, and 4 degrees C for the carbonmonoxy ones. There was no significant difference in stability between the met and carbonmonoxy crosslinked proteins. In order of increasing stability the samples were: met Hb S less than met Hb A less than CO Hb S less than CO Hb A = CN-met Hb A less than met XL-Hb S = CO XL-Hb S less than met XL-Hb A = CO XL-Hb A less than CN-met XL-Hb A. The slight decrease in the stability of Hb S (beta 6 Glu----Val) compared to Hb A can be explained by the replacement of an external ionic group by a hydrophobic residue in Hb S. In mixtures of crosslinked and normal Hb A, the Tm of the uncrosslinked material was slightly increased by the presence of the more stable crosslinked hemoglobin. The effects of both crosslinking and cyanide or carbon monoxide binding can be explained by Le Chatelier's principle since both would favor the native form of the protein.  相似文献   

14.
Hemoglobin-based oxygen carriers (HBOCs) are candidates for use as blood substitutes and resuscitation fluids. We determined that HBOCs of specific types differ in their ability to generate or interact with free radicals. The differences do not correlate with oxygen affinity. Detailed comparisons with unmodified human hemoglobin, HbA0, were carried out with two cross-linked derivatives: HbA-FMDA, produced by the reaction of human oxyhemoglobin with fumaryl monodibromoaspirin, and HbA-DBBF, produced by the reaction of human deoxyhemoglobin with bis(3,5-dibromosalicyl) fumarate. Both derivatives had lower oxygen affinity than unmodified HbA0. As previously reported, exposure of oxyhemoglobin to H2O2 causes generation of free radicals capable of generating formaldehyde from dimethyl sulfoxide. Relative to the reaction catalyzed by 50 microM HbA (18.0 +/- 3.5 nmol/30 min/ml), the formaldehyde formation was roughly 70% for HbA-DBBF and 50% for HbA-FMDA under comparable conditions. More profound differences are exhibited at lower hemoglobin concentrations. Spectral changes of the HBOCs during the reaction differ qualitatively and occur at different rates. The HBOCs also differ in rates of hemoglobin-catalyzed NADPH oxidation and aniline hydroxylation, reactions mediated by reactive oxygen species. These results show that stereochemical differences brought about by chemical cross-linking alter the ability of HBOCs to generate radicals and to react with activated oxygen species. These studies also show that the ability of hemoglobin to produce activated species of oxygen can be enhanced or suppressed independently of oxygen affinity.  相似文献   

15.
Previous studies point to the acidic amino-terminal segment of band 3, the anion transport protein of the red cell, as the common binding site for hemoglobin and several of the glycolytic enzymes to the erythrocyte membrane. We now report on the interaction of hemoglobin with the synthetic peptide AcM-E-E-L-Q-D-D-Y-E-D-E, corresponding to the first 11 residues of band 3, and with the entire 43,000-Da cytoplasmic domain of the protein. In the presence of increasing concentrations of the peptide, the oxygen binding curve for hemoglobin is shifted progressively to the right, indicating that the peptide binds preferentially to deoxyhemoglobin. The dissociation constant for the deoxyhemoglobin-peptide complex at pH 7.2 in the presence of 100 mM NaCl is 0.31 mM. X-ray crystallographic studies were carried out to determine the exact mode of binding of the peptide to deoxyhemoglobin. The difference electron density map of the deoxyhemoglobin-peptide complex at 5 A resolution showed that the binding site extends deep (approximately 18 A) into the central cavity between the beta chains, along the dyad symmetry axis, and includes Arg 104 beta 1 and Arg 104 beta 2 as well as most of the basic residues within the 2,3-diphosphoglycerate binding site. The peptide appears to have an extended conformation with only 5 to 7 of the 11 residues in contact with hemoglobin. In agreement with the crystallographic studies, binding of the peptide to deoxyhemoglobin was blocked by cross-linking the beta chains at the entrance to the central cavity. Oxygen equilibrium studies showed that the isolated cytoplasmic fragment of band 3 also binds preferentially to deoxyhemoglobin. The binding of the 43,000-Da fragment to hemoglobin was inhibited in the cross-linked derivative indicating that the acidic amino-terminal residues in the intact cytoplasmic domain also bind within the central cavity of the hemoglobin tetramer.  相似文献   

16.
Previous studies have shown that pteroylheptaglutamate (PteGlu7) can form a 1:1 complex with deoxyhemoglobin. The solution and crystallographic studies reported in this paper delineate the nature of the PteGlu7 binding site. We find that the three structural elements of PteGlu7 (the pteridine moiety, the p-aminobenzoyl portion, and the glutamate groups) each contribute to the binding energy by interacting with residues in the central cavity between the beta subunits and with residues at the alpha 1 beta 1 interface. Identification of the 2,3-diphosphoglycerate (DPG) binding site as part of the PteGlu7 binding site was accomplished in two ways; first by the demonstration of reduced PteGlu7 binding to hemoglobin selectively modified by pyridoxylation at this site, and second by the finding that DPG and PteGlu7 bind to deoxyhemoglobin in a competitive manner. In addition, since analogs of PteGlu7 in which the pteridine moiety is modified display reduced binding, it can be concluded that the pteridine group also contributes significantly to the binding energy. The crystallographic studies are completely consistent with the results determined in solution. A difference electron density image at 4.3 A resolution shows that the pteridine and p-aminobenzoyl groups are nestled against an interior edge of the alpha 1 beta 1 interface with the pteridine ring interacting with Phe 36 alpha 1 and the p-aminobenzoyl group positioned against a portion of the H helix between residues Lys 132 beta 1 and Ala 135 beta 1. The difference density for the glutamate residues is less well resolved (for reasons described in the text), but it is clear that some of the carboxylate side chains must interact with residues at the DPG binding site.  相似文献   

17.
The synthesis of N-(2,4-diphosphobenzyl)-1-amino-5-naphthalenesulfonic acid (DIPANS) is described. It entails the synthesis of 2,4-diphosphobenzaldehyde from the action of POCl3 on 2,4-dihydroxybenzaldehyde. This is followed by coupling of the 2,4-diphosphobenzaldehyde to 1-amino-5-naphthalenesulfonic acid. Subsequent reduction with NaBH4 yields the desired product. The DIPANS exhibits an excitation maximum at 337 nm and a fluorescence emission maximum at 504 nm. This dye is quantitatively displaced by inositol hexaphosphate and is an effective analogus of diphosphoglyceric acid (DPG), possessing a KD at pH 7.0 in 0.05 M [bis(2-hydroxyethyl)amino]tris(hydroxymethyl)methane (bis-Tris) plus 0.1 M chloride of 6.88 microgram, with 1.0 molecule bound/hemoglobin tetramer. Like DPG its binding to deoxyhemoglobin decreases with increasing pH; in the presence of 0.1 M chloride it binds 0.031 times as tightly to CO hemoglobin and it yields a value for free energy coupling of 2.0 kcal/mol. The presence of 1 mM DIPANS decreases the affinity of hemoglobin for oxygen in the absence of salt from p1/2 of 0.8 mm Hg to 12.4 mm Hg. Using DPG as a competitor of DIPANS binding, a dissociation constant of 11.4 micrometer was calculated for DPG binding to deoxy-Hb at pH 7.0 in the presence of 0.05 M bis-Tris and 0.1 M chloride.  相似文献   

18.
Methyl acetyl phosphate binds to the 2,3-diphosphoglycerate (2,3-DPG) binding site of hemoglobin and selectively acetylates three amino groups at or near that site. The subsequent binding of 2,3-DPG is thus impeded. When intact sickle cells are exposed to methyl acetyl phosphate, their abnormally high density under anaerobic conditions is reduced to the density range of oxygenated, nonsickling erythrocytes. This change is probably due to a combination of direct and indirect effects induced by the specific acetylation. The direct effect is on the solubility of deoxyhemoglobin S, which is increased from 17 g/dL for unmodified hemoglobin S to 22 g/dL for acetylated hemoglobin S at pH 6.8. Acetylated hemoglobin S does not gel at pH 7.4, up to a concentration of 32 g/dL. The indirect effect could be due to the decreased binding of 2,3-DPG to deoxyhemoglobin S within the sickle erythrocyte, thus hindering the conversion of oxyhemoglobin S to the gelling form, deoxyhemoglobin S.  相似文献   

19.
Park SY  Shibayama N  Hiraki T  Tame JR 《Biochemistry》2004,43(27):8711-8717
A number of ligand binding studies of human adult hemoglobin (HbA) cross-linked between Lys 82beta(1) and Lys 82beta(2) with bis(3,5-dibromosalicyl)fumarate have been reported. The oxygen binding properties of native HbA, including the cooperativity and Bohr effect, are not substantially changed by the modification, provided care is taken to remove electrophoretically silent impurities arising from side reactions. We have refined the high-resolution structure of this modified Hb and found it adopts the T state when crystallized in the absence of heme ligands, contrary to a previously published structure. These results suggest the slightly altered crystal form determined previously may be due to unremoved side products of the cross-linking reaction with high oxygen affinity. Two nickel-substituted Hbs cross-linked in the same way have also been crystallized in the presence of carbon monoxide, which binds only to the ferrous heme. In the case of the nickel-substituted alpha subunit, the absence of a covalent link between the central metal of the heme and the proximal histidine leads to a new conformation of the histidine stabilized by a water molecule. This structure may mimic that of partially NO-liganded species of HbA; however, overall, the changes are highly localized, and both doubly ligated species are in the T conformation.  相似文献   

20.
Comparative data on quaternary structure, cooperativity, Bohr effect and regulation by organic phosphates are reviewed for vertebrate hemoglobins. A phylogeny of hemoglobin function in the vertebrates is deduced. It is proposed that from the monomeric hemoglobin of the common ancestor of vertebrates, a deoxy dimer, as seen in the lamprey, could have originated with a single amino acid substitution. The deoxy dimer has a Bohr effect, cooperativity and a reduced oxygen affinity compared to the monomer. One, or two, additional amino acid substitutions could have resulted in the origin of a tetrameric deoxy hemoglobin which dissociated to dimers on oxygenation. Gene duplication, giving incipient alpha and beta genes, probably preceded the origin of a tetrameric oxyhemoglobin. The origin of an organic phosphate binding site on the tetrameric hemoglobin of an early fish required only one, or two, amino acid substitutions. ATP was the first organic phosphate regulator of hemoglobin function. The binding of ATP by hemoglobin may have caused the original elevation in the concentration of ATP in the red blood cells by relieving end product inhibition of ATP synthesis. The switch from regulation of hemoglobin function by ATP to regulation by DPG may have been a consequence of the curtailment of oxidative phosphorylation in the red blood cell. The basic mechanisms by which ATP and DPG concentrations can respond to strss on the oxygen transport system were present before the origin of an organic phosphate binding site on hemoglobin. A switch from ATP regulation to IP5 regulation occurred in the common ancestor of birds.  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