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 312 毫秒
1.
2.
《Biophysical journal》2022,121(14):2767-2780
Hemoglobins M (Hbs M) are human hemoglobin variants in which either the α or β subunit contains a ferric heme in the α2β2 tetramer. Though the ferric subunit cannot bind O2, it regulates O2 affinity of its counterpart ferrous subunit. We have investigated resonance Raman spectra of two Hbs, M Iwate (α87His → tyrosine [Tyr]) and M Boston (α58His → Tyr), having tyrosine as a heme axial ligand at proximal and distal positions, respectively, that exhibit unassigned resonance Raman bands arising from ferric (not ferrous) hemes at 899 and 876 cm-1. Our quantum chemical calculations using density functional theory on Fe-porphyrin models with p-cresol and/or 4-methylimidazole showed that the unassigned bands correspond to the breathing-like modes of Fe3+-bound Tyr and are sensitive to the Fe-O-C(Tyr) angle. Based on the frequencies of the Raman bands, the Fe-O-C(Tyr) angles of Hbs M Iwate and M Boston were predicted to be 153.5° and 129.2°, respectively. Consistent with this prediction, x-ray crystallographic analysis showed that the Fe-O-C(Tyr) angles of Hbs M Iwate and M Boston in the T quaternary structure were 153.6° and 134.6°, respectively. It also showed a similar Fe-O bond length (1.96 and 1.97 Å) and different tilting angles.  相似文献   

3.
Solvent-proton longitudinal magnetic relaxation rates as dependent on temperature were measured for human (H)/canine (C) valency hybrids of the type {αH(III)βC(II)}2 and {αC(II)βH(III)}2. The two metheme irons in the human methemoglobin chains induce quite different proton magnetic relaxation (pmr) rates reflecting a tighter β-heme-pocket compared to the α subunit. Both heme-pockets appear to be loosened in the presence of inositol hexaphosphate (IHP) although this allosteric effector binds only to the β chains, the binding assumed to be the same for canine as for human hemoglobin. The subunit nonequivalence is retained also in the T-quaternary state induced by IHP. In the species hybrids the pmr rates due to the metheme iron are sensitive to the valency (ligand) state, which was either CO or H2O in the partner half of the hybrid. All results show very clearly the interrelationship of the tertiary (protomer) structure with the quaternary (oligomer) structure in hemoglobin.  相似文献   

4.
The improved methods for the preparation of valency hybrid hemoglobins, (α3+β2+)2 and (α2+β3+)2 were presented. The (α3+β2+)2 valency hybrid was separated from the solutions of partially reduced methemoglobin with ascorbic acid, by using CM 32 column chromatography. The (α2+β3+)2 valency hybrid was also isolated from hemoglobin solutions, which were partially oxidized with ferricyanide, by chromatography on CM 32 column. These valency hybrid hemoglobins were found to be single on isoelectric focusing electrophoresis. Present procedures are very simple and are suitable for the bulk preparation of (α3+β2+)2 and (α2+β3+)2 valency hybrids.  相似文献   

5.
Haemoglobin valency hybrids have been further investigated with a view to evaluating evidence for α-β interactions. Comparison of equilibrium oxygen binding of the compounds αIIIH2OβII, αIIIFβII, αIIIN3βII and αIIICNβII show that the derivatives possessing the α chain in the (low spin) oxy conformation possess higher oxygen affinity than those possessing the same chain in (high spin) deoxy conformation. On the other hand, equilibrium titration of the aquo derivative by fluoride and azide showed a higher azide affinity and a slightly lower fluoride affinity for αIIIH2OβIICO compared to the deoxy form αIIIH2OβII. Essentially the same pattern of relations were obtained for the different forms of αIIβII also.Electron paramagnetic resonance spectra of the hybrids at pH 6 showed no change of the g value of 5.85 absorption of the ferri chain on change of spin state of the partner ferro chain. However, this was no longer the case at pH 9; the spectra of the alkaline form gave evidence of α-β interaction for the αIIIβII hybrid only, but not for αIIβIII. The electron paramagnetic resonance results suggest that α-β interaction in the hybrids may operate in the β → α direction only.The equilibrium and spectroscopic data are discussed in the general context of haem-haem interaction.  相似文献   

6.
Oxygen equilibrium determinations with “unsymmetrical” MetHb/Hb hybrids derived from human hemoglobins A and S are reported. All four of the possible hybrids have higher oxygen affinity than the parent hemoglobins. The α2Metβ2S hybrid has a lower oxygen affinity than that of α2Metβ2S. However, both the βMet hybrids have similar oxygen affinity. The Bohr value of α2Metβ2S is more negative than that of α2Metβ2A while the βMet hybrids appear to have almost identical Bohr values. These findings favor the view that α and β chains in hemoglobin A have different conformations and indicate that hemoglobin S has a β-chain conformation different from that of β-chain of hemoglobin A. This difference is probably carried into the oxygenation properties of the α-chain in such a way as to be reflected only when the β chain is oxidized.  相似文献   

7.
Mo(CO)4(LL) complexes, where LL = polypyridyl ligands such as 2,2′-bipyridine and 1,10-phenanthroline, undergo quasi-reversible, one-electron oxidations in methylene chloride yielding the corresponding radical cations, [Mo(CO)4(LL)]+. These electrogenerated species undergo rapid ligand substitution in the presence of acetonitrile, yielding [Mo(CO)3(LL)(CH3CN)]+; rate constants for these substitutions were measured using chronocoulometry and were found to be influenced by the steric and electronic properties of the polypyridyl ligands. [Mo(CO)3(LL)(CH3CN)]+ radical cations, which could also be generated by reversible oxidation of Mo(CO)3(LL)(CH3CN) in acetonitrile, can be irreversibly oxidized yielding [Mo(CO)3(LL)(CH3CN)2]2+ after coordination by an additional acetonitrile. Infrared spectroelectrochemical experiments indicate the radical cations undergo ligand-induced net disproportionations that follow first-order kinetics in acetonitrile, ultimately yielding the corresponding Mo(CO)4(LL) and [Mo(CO)2(LL)(CH3CN)3]2+ species. Rate constants for the net disproportionation of [Mo(CO)3(LL)(CH3CN)]+ and the carbonyl substitution reaction of [Mo(CO)3(LL)(CH3CN)2]2+ were measured. Thin-layer bulk oxidation studies also provided infrared characterization data of [Mo(CO)4(ncp)]+ (ncp = neocuproine), [Mo(CO)3(LL)(CH3CN)]+, [Mo(CO)3(LL)(CH3CN)2]2+ and [Mo(CO)2(LL)(CH3CN)3]2+ complexes.  相似文献   

8.
The improved methods for the preparation of valency hybrid hemoglobins, (α3+β2+)2 and (α2+β3+)2 were presented. The (α3+β2+)2 valency hybrid was separated from the solutions of partially reduced methemoglobin with ascorbic acid, by using CM 32 column chromatography. The (α2+β3+)2 valency hybrid was also isolated from hemoglobin solutions, which were partially oxidized with ferricyanide, by chromatography on CM 32 column. These valency hybrid hemoglobins were found to be single on isoelectric focusing electrophoresis. Present procedures are very simple and are suitable for the bulk preparation of (α3+β2+)2 and (α2+β3+)2 valency hybrids.  相似文献   

9.
The Δ-α-and Λ-β- forms of [Co(edda)CO3]? have been isolated as crystalline diastereoisomeric salts of the cation Δ-[Co(en)2(ox)]+. When each salt is dissolved in 1.0 F aqueous Na2CO3 at 25°C an isomeric equilibration process occurs between these species in which the Δ-α form predominates before the eventual racemisation of both anionic species. On proceeding to equilibrium the Λ-β-[Co(edda) CO3]? isomer is observed to invert its absolute configuration at the central metal ion in converting to the Δ-α stereoisomer. This is the first reported example of inversion-isomerisation involving the β → α transformation in a chiral complex containing a linear tetradentate.  相似文献   

10.
The proton magnetic resonance spectra of the dihydronicotinamide ring of αNADH3 and the nicotinamide ring of αNAD+ are reported and the proton absorptions assigned. The absolute assignment of the C4 methylene protons of αNADH is based on the generation of specifically deuterium-labeled (pro-S) B-deuterio-αNADH from enzymatically prepared B-deuterio-βNADH. The C4 proton absorption of αNAD+ is assigned by oxidation of B-deuterio-αNADH by the A specific, yeast alcohol dehydrogenase to yield 4-deuterio-αNAD+.The epimerization of either αNADH or βNADH yields an equilibrium ratio of approximately 9:1 βNADH to αNADH. The rate of epimerization of αNADH to βNADH at 38 °C in 0.05, pH 7.5, phosphate buffer is 3.1 × 10?3 min?1, corresponding to a half-life of 4 hr. Four related dehydrogenases, yeast and horse liver alcohol dehydrogenase and chicken M4 and H4 lactate dehydrogenase, are shown to oxidize αNADH to αNAD+ at rates three to four orders of magnitude slower than for βNADH. By using specifically labeled B-deuterio-αNADH the enzymatic oxidation by yeast alcohol dehydrogenase has been shown to occur with the identical stereospecificity as the oxidation of βNADH. The nonenzymatic epimerization of αNADH to βNADH and the enzymatic oxidation αNADH are discussed as a possible source of αNAD+in vivo.  相似文献   

11.
Cyclic voltammetry in acetonitrile has been used to study the redox properties of the compound [Cu-(BBDH)Cl] Cl under various conditions. The nature of the species in solution was studied with the aid of ligand-field spectra and ESR spectra. It appears that in CH3CN solution two Cu(II) species are predominating, viz. [Cu(BBDH)(CH3CN)x]2+ and Cu(BBDH)-Cl+. Minor species are dimeric in nature, such as [(BBDH)CuCl2Cu(BBDH)]2+, as easily seen from ESR. The relative amount of the two major species was varied using added LiCl, allowing us to determine the nature of both species. The redox potential of the species [Cu(BBDH)Cl]+ appeared to be 0.62 V (against a normal hydrogen electrode), which is very high for a Cu(II) compound and in the same area as found for the blue copper proteins. Re-oxidation of Cu(BBDH)+ in the presence of LiCl shows that Cl slowly recoordinates after reoxidation.  相似文献   

12.
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   

13.
Observation of allosteric transition in hemoglobin   总被引:6,自引:0,他引:6  
Two conclusions have been drawn from NMR studies of mixed state hemoglobins. First the α and β subunits in hemoglobin are not equivalent in their conformational properties. Second the mixed state hemoglobin (αIIICN βII)2 can take two different quaternary structures without changing the degree of ligation. One of the two structures is similar to that of deoxyhemoglobin and the other to that of oxyhemoglobin.  相似文献   

14.
Specific salt effects were studied on the quenching reaction of excited [Ru(NN)3]2+ (NN=2,2′-bipyridine(bpy), 1,10-phenanthrorine(phen)) and [Cr(bpy)3]3+ by [Cr(CN)6]3−, [Fe(CN)6]3− and [Ni(CN)4]2− in aqueous solutions as a function of alkali metal ions which were added for adjustment of ionic strength. The quenching rate constants in [Ru(NN)3]2+-[Cr(CN)6]3− and [Cr(bpy)3]3+-[Cr(CN)6]3− systems are changed by the cations as Li+>Na+>K+≈Rb+≈Cs+. On the other hand, the rate constants in [Ru(NN)3]2+-[Fe(CN)6]3− and [Ru(NN)3]2+-[Ni(CN)4]2− systems, which are diffusion-controlled reactions, are not varied by the alkali metal cations. The obtained order (Li+>Na+>K+≈Rb+≈Cs+) of the quenching rate constant is quite different from salt effects, Li+<Na+<K+<Rb+<Cs+, which have been obtained in the electron transfer reactions between complex anions.  相似文献   

15.
The previously reported complex [Ru(ttpy)(CN)3] [ttpy = 4′(p-tolyl)-2,2′:6′,2″-terpyridine] is conveniently synthesised by reaction of ttpy with Ru(dmso)4Cl2 to give [Ru(ttpy)(dmso)Cl2], which reacts in turn with KCN in aqueous ethanol to afford [Ru(ttpy)(CN)3] which was isolated and crystallographically characterised as both its (PPN)+ and K+ salts. The K+ salt contains clusters containing three complex anions and three K+ cations connected by end-on and side-on cyanide ligation to the K+ ions. The solution speciation behaviour of [Ru(ttpy)(CN)3] was investigated with both Zn2+ and K+ salts in MeCN, a solvent sufficiently non-competitive to allow the added metal cations to associate with the complex anion via the externally-directed cyanide lone pairs. UV-Vis spectroscopic titration of (PPN)[Ru(ttpy)(CN)3] with Zn(ClO4)2 showed a blue shift of 2900 cm−1 in the 1MLCT absorption manifold due to the ‘metallochromism’ effect; a series of distinct binding events could be discerned corresponding to formation of 4:1, 1:1 and then 1:3 anion:cation adducts, all with high formation constants, as the titration proceeded. In contrast titration of (PPN)[Ru(ttpy)(CN)3] with the more weakly Lewis-acidic KPF6 resulted in a much smaller blue-shift of the 1MLCT absorptions, and the titration data corresponded to formation of 1:1 and then 2:1 cation:anion adducts with weaker stepwise association constants of the order of 104 and then 103 M−1. Although association of [Ru(ttpy)(CN)3] resulted in a blue-shift of the 1MLCT absorptions, the luminescence was steadily quenched, as raising the 3MLCT level makes radiationless decay via a low-lying 3MC state possible.  相似文献   

16.
The translation of rabbit α globin mRNA in a Krebs II ascites cellfree system was more dependent upon the K+ concentration than rabbit β globin mRNA. The optimal KCl concentration was approximately 70 mM for the synthesis of the α chain and between 80 and 90 mM for that of the β chain. With CH3 CO2K the optimum concentration for α chain synthesis was also 70 mM but the optimum for the β chain synthesis was not sharp any more and ranged from 70 mM to over 110 mM. In the range of the optimal Mg2+ concentration for the α and β globin chain synthesis the αβ ratio decreased when the Mg2+ concentration increased. In the presence of DTT and EDTA the optimal KCl concentration for both α and β globin chain synthesis decreased.  相似文献   

17.
18.
The formation constant, log β4 = 62.3 for [Pd(CN)4]2− is reported at 25 °C in 0.1 M NaClO4. This value of log β4 was determined using a competition reaction, monitored using UV-Vis spectroscopy and 1H NMR. The competition reaction used was with the tetraamine ligand 2,3,2-tet(1,4,8,11-tetraazaundecane), for which log K1 = 47.8 at 25 °C in 0.1 M NaClO4 was determined by competition with thiocyanate, as described by earlier workers (Q.Y. Yan, G. Anderegg, Inorg. Chim. Acta 105 (1985) 121.). Also reported is a value of log β4 for the [Pd(SCN)4]2− ion of 27.2 in 0.1 M NaClO4, determined by competition with 2,2,2-tet. Measurement of log K1 for cyclam with Pd(II) was attempted using a competition reaction with cyanide, combined with the very high value of log β4 for [Pd(CN)4]2− measured here. It appeared that the equilibrium being followed was actually [Pd(cyclam)]2+ + 2CN ? [Pd(cyclam)(CN)2], for which a constant of log K = 5.2 was obtained. 1H NMR and IR studies suggested that the complex [Pd(cyclam)(CN)2] was prone to oxidation to Pd(IV), followed by disproportionation to [Pd(cyclam)]2+ and, presumably, (CN)2. The very high value of log β4 for [Pd(CN)4]2− found here appears to be the highest formation constant known for any metal ion.  相似文献   

19.
Infrared spectra of 15N-enriched preparations of the soluble cytoplasmic NAD+-reducing [NiFe]-hydrogenase from Ralstonia eutropha are presented. These spectra, together with chemical analyses, show that the Ni-Fe active site contains four cyanide groups and one carbon monoxide molecule. It is proposed that the active site is a (RS)2(CN)Ni(-RS)2Fe(CN)3(CO) centre (R=Cys) and that H2 activation solely takes place on nickel. One of the two FMN groups (FMN-a) in the enzyme can be reversibly released upon reduction of the enzyme. It is now reported that at longer times also one of the cyanide groups, the one proposed to be bound to the nickel atom, could be removed from the enzyme. This process was irreversible and induced the inhibition of the enzyme activity by oxygen; the enzyme remained insensitive to carbon monoxide. The Ni-Fe active site was EPR undetectable under all conditions tested. It is concluded that the Ni-bound cyanide group is responsible for the oxygen insensitivity of the enzyme.Abbreviations BV benzyl viologen - DCIP 2,6-dichlorophenol-indophenol - EXAFS extended X-ray absorption fine structure - FTIR Fourier transform infrared - MV methyl viologen - SH soluble NAD+-reducing hydrogenase - XAS X-ray absorption spectroscopy  相似文献   

20.
In this study, the fluorescence spectra of sarafloxacin (SAR) under different pH conditions were investigated to determine the structural changes due to protonation that result from change in pH. At pH < 1.02, SAR exists in the H3L2+ form for which the maximum fluorescence emission wavelength was about 455 nm. At pH 1.87–4.94, SAR exists in the H2L+ form in which H3L2+ loses one proton in the nitrogen molecule at the 1‐position in the quinoline ring. Fluorescence intensity was strong and steady and the maximum emission wavelength was 458 nm. At pH 7.14–9.30, the maximum emission wavelengths were gradually blue shifted to 430 nm with increase in pH, here SAR exists in the form of a bipolar ion HL in which H2L+ loses a carboxyl group proton. At pH > 11.6, HL transforms into anionic L? in which HL loses one proton from the piperazine ring, leading to a decrease in fluorescence intensity, and the maximum emission wavelength was red shifted to approximately 466 nm. The two‐step dissociation constant pKa for SAR was calculated, pK a1 was 6.06 ± 0.37 and pK a2 for SAR was 10.53 ± 0.19. In a pH 3.62 buffer solution with quinine sulfate as the reference, the fluorescence quantum yield of SAR at the maximum excitation wavelength of 276 nm was 0.09.  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