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 15 毫秒
1.
Asymmetrical hybrid hemoglobins formed from mixtures of oxyhemoglobins S and F and A and F were separated by high-performance liquid chromatography on a 4.6 X 250 mm wide-pore polyethyleneimine-silica gel column under anaerobic conditions. The resulting HPLC chromatogram showed three peaks, with the middle peak representing the hybrid hemoglobin. The areas of these three peaks were quantified and the amount of hybrids formed was less than that predicted theoretically. We found that the deviation was due to the equilibrium constant of the FS hybrid hemoglobin differing from that of the parent hemoglobins. In this report, we introduce the anaerobic recycle ion-exchange HPLC method to determine the rate of dissociation of AS and FS hybrid hemoglobins at constant pH buffer conditions. The results obtained by this method demonstrate that FS hybrid hemoglobin is more unstable than AS hybrid hemoglobin. The free energy of association for asymmetrical hybrids containing hemoglobin F is approximately 0.6 Kcal/mol greater than that of the symmetrical parent hemoglobins.  相似文献   

2.
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   

3.
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   

4.
The hemoglobins of Spirocamallanus cricotus, a reddish-colored, camallanid nematode, and its Atlantic croacker fish host, Micropogonias undulatus, were characterized with spectrophotometry and isoelectric focusing. Hemoglobin from female parasites' perienteric fluid and homogenized male parasites gave Spectrophotometric peaks at 412, 539, and 575 nm, whereas female worms drained of perienteric fluid and homogenized differed by having a Soret peak of 408 nm. Changing the ionic strength of the buffer from 0.1 to 0.01 M shifted the Soret peak to 406 nm for the female parasites' perienteric fluid and ground male parasites and 404 nm for homogenized female parasites. In all cases, the β band had a higher absorption than the α band suggesting a high O2 affinity for the parasite hemoglobin. Host hemoglobin had peaks of 406, 437, and 577 nm. Isoelectric focusing not only confirmed the Spectrophotometric evidence that host and parasite hemoglobins differed, but also showed that the parasite's analyzed hemoglobin fractions differed from one another by having different isoelectric points.  相似文献   

5.
A simpler, more economical technique than previously reported, that of conventional polyacrylamide gel electrophoresis alone, is described for the detection of asymmetric hemoglobin hybrids of the forms alphaXalphaYbeta2 and alpha2betaXbetaY when bloods from individuals with alpha and beta chain variants were examined. The presence of alpha chain variant hybrids, never before reported, is further evidence that hybrid formation is a more widespread phenomenon than has previously been thought of. Hybrids were found in artificial mixtures of hemoglobins and more importantly, are also reported here for the first time in bloods of individuals heterozygous for hemoglobin variants. These hybrid tetramers were as stable as the parent hemoglobins when examined under anaerobic conditions. The involvement of HbF in the formation of hybrids of the type alpha2betagamma is reported, and an analysis of the possible role of these as well as alpha2betaAbetaS hybrids in the sickling process is presented.  相似文献   

6.
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   

7.
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   

8.
  • 1.1. Analysis of the Soret spectra of hemoglobins A, S and F has been used to determine the extent of heme exposure and release from these hemoglobins in the presence of several solvent perturbants.
  • 2.2. Oxyhemoglobin S unfolding in the presence of either urea or propyl urea resulted in greater heme exposure and release than either oxyhemoglobins A or F.
  • 3.3. Methemoglobin formation resulted in lower denaturation midpoints for each hemoglobin compared to the reduced oxyhemoglobin state; methemoglobin F had the lowest denaturation midpoint under isothermal denaturing conditions.
  • 4.4. Rate of heme exposure was greater for oxyhemoglobin S than oxyhemoglobin A in the presence of 200 μM the anionic detergent sodium dodecyl sulfate.
  • 5.5. Evidence for increased levels of heme release in hemoglobin S may be related to the greater tendency of sickled red cell membranes to undergo lipid oxidation.
  相似文献   

9.
Binding of triethyltin to the cat hemoglobins (HbA and HbB) results in the “masking” of two of the freely reactive sulfhydryl groups (SH) within the hemoglobin tetramer. That the “masked” SH groups occur in position 13α of each α-subunit was demonstrated by the lack of labeling of cysteine 13α with [14C]N-ethylmaleimide when triethyltin is present. Studies with cat-human hybrid hemoglobins indicate that the α-subunit of the cat hemoglobins alone is involved in the formation of a complex with triethyltin. Using available data on the primary as well as three dimensional structures of animal hemoglobins, it is suggested the cysteine 13α and histidine 20α serve as axial ligands in the formation of a pentacoordinate triethyltin cat hemoglobin complex. The binding of triethyltin results in an increase in the oxygen affinity of the two cat hemoglobins.  相似文献   

10.
A new class of hybrid nitric oxide-releasing anti-inflammatory (AI) ester prodrugs (NONO-coxibs) wherein an O2-acetoxymethyl-1-(N-ethyl-N-methylamino)diazen-1-ium-1,2-diolate (13ab), or O2-acetoxymethyl-1-(2-methylpyrrolidin-1-yl)diazen-1-ium-1,2-diolate (16ab), NO-donor moiety was covalently coupled to the COOH group of 5-(4-carboxymethylphenyl)-1-(4-methane(amino)sulfonylphenyl)-3-trifluoromethyl-1H-pyrazole (11ab) was synthesized. The percentage of NO released from these diazen-1-ium-1,2-diolates was significantly higher (59.6–74.6% of the theoretical maximal release of 2 molecules of NO/molecule of the parent hybrid ester prodrug) upon incubation in the presence of rat serum, relative to incubation with phosphate buffer (PBS) at pH 7.4 (5.0–7.2% range). These incubation studies suggest that both NO and the AI compound would be released from the parent NONO-coxib upon in vivo cleavage by non-specific serum esterases. All compounds were weak inhibitors of the COX-1 isozyme (IC50 = 8.1–65.2 μM range) and modest inhibitors of the COX-2 isozyme (IC50 = 0.9–4.6 μM range). The most potent parent aminosulfonyl compound 11b exhibited AI activity that was about sixfold greater than that for aspirin and threefold greater than that for ibuprofen. The ester prodrugs 13b, 16b exhibited similar AI activity to that exhibited by the more potent parent acid 11b when the same oral μmol/kg dose was administered. These studies indicate hybrid ester AI/NO donor prodrugs of this type (NONO-coxibs) constitute a plausible drug design concept targeted toward the development of selective COX-2 inhibitory AI drugs that are devoid of adverse cardiovascular effects.  相似文献   

11.
Lepore hemoglobins result from crossovers between normal beta and delta chain genes. Structural investigation of two newly discovered examples of Lepore hemoglobins revealed one of them to be structurally identical to hemoglobin Lepore Hollandia α2Aδ22 -x- β50, a rarely occurring Lepore variant, while the second had the structure of hemoglobin Lepore Boston α2Aδ87 -x- β116. Studies of the equilibrium and kinetic properties of the liganding reactions of these two Lepore hemoglobins, which differ only in three amino acid residues, and comparison of these with the known properties of hemoglobin A1 (α2β2) and hemoglobin A2 (α2δ2) have been carried out. A high value of n, the Hill coefficient, indicating normal heme-heme interaction, was observed in each hemoglobin along with a normal Bohr effect. However, a slight but definite increase in oxygen affinity was observed for each Lepore hemoglobin. Furthermore, kinetic studies indicated a slight but consistently increased rate of ligand combination and a somewhat decreased rate of oxygen dissociation for hemoglobins Lepore Hollandia and Lepore Boston at pH 7 and 20 °C. Apparently, the higher oxygen affinity of these Lepore hemoglobins over those of the normal hemoglobins A1 and A2 reflects changes of sequence that are common to both types of hemoglobin Lepore.  相似文献   

12.
《BBA》2023,1864(2):148960
Protochlorophyllide(PChlide)-a and its 8-vinylated analog, divinyl(DV)-PChlide-a, are common and essential intermediates in the biosynthesis of all naturally occurring chlorophyll (Chl) pigments. These porphyrinoid-type pigments have a single optically active (asymmetric) carbon atom at the 132-position, so their stereoisomers are (132R)- and (132S)-enantiomers. The former and latter are called (DV-)PChlide-a and (DV-)PChlide-a′, respectively. In this study, chiral-phase HPLC separation of enantiomeric (DV-)PChlides-a/a′ was demonstrated. The (132R)-enantiomeric PChlide-a was eluted more slowly than the corresponding (132S)-enantiomeric PChlide-a′ under the present HPLC conditions. On the other hand, the elution order of (132R)-DV-PChlide-a and (132S)-DV-PChlide-a′ was reverse to that of PChlides-a/a′. After the separation of each enantiomer by the chiral-phase HPLC, the stereoisomeric configuration at the 132-position was characterized by means of circular dichroism spectroscopy. The present chiral-phase HPLC method enables us to evaluate optical purities of (DV-)PChlide-a species. For example, PChlide-a and/or DV-PChlide-a extracted from the spent medium and harvested cells of cultured purple photosynthetic bacterial mutants, the former of which has been often used as the source of (DV-)PChlide-a substrates for enzymatic reactions, were revealed to be mostly racemized, giving enantiomeric mixtures of (DV-)PChlides-a/a′.  相似文献   

13.
The ferric hemes of valence hybrid hemoglobins combine with imidazole in a manner analogous with the hemes of methemoglobin. Equilibrium studies show that imidazole binding to methemoglobin is minimally described by the sum of two independent processes (K1 = 200 M?1 and K2 = 37 M?1), both of which contribute equally to the observed difference spectrum. Using valance hybrid hemoglobins, which show single binding processes under similar conditions, it is possible to identify the high affinity sites in methemoglobin with the α chains and the low affinity sites with the β chains.Kinetic studies show that the valance hybrid hemoglobins react in a single exponential fashion with imidazole in contrast with methemoglobin which shows a biphasic reaction (k1 = 85 M?1 sec?1k2 = 25 M?1 sec?1). A comparison of the rates of reaction of the hybrids allows the assignment of the fast phase in methemoglobin to the β chains and the slow phase to the α chains.The heterogeneity of the imidazole reaction with methemoglobin occurs over the pH range 5.5–9.5 within which two ionization processes are discernable at pH 6.9 and 7.5.  相似文献   

14.
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   

15.
This paper shows that the “second Emerson effect”1 exists not only in photosynthesis, but also in the quinone reduction (Hill reaction), in Chlorella pyrenoidosa and Anacystis nidulans. The peaks at 650 mμ, 600 mμ, 560 mμ, 520 mμ, and 480 mμ, observed in the action spectrum of this effect in the Hill reaction in Chorella, are attributable to chlorophyll b; the occurrence of an additional peak at 670 mμ, 620 mμ, and of two (or three) peaks in the blueviolet region suggests that (at least) one form of chlorophyll a contributes to it. In analogy to suggestions made previously in the interpretation of the Emerson effect in photosynthesis, these results are taken as indicating that excitation by light preferentially absorbed by one (or two) forms of chlorophyll a (Chl a 690 + 700), needs support by simultaneous absorption of light in another form of chlorophyll a (Chl a 670)—directly or via energy transfer from chlorophyll b—in order to produce the Hill reaction with its full quantum yield. In Anacystis, the participation of phycocyanin in the Emerson effect in the Hill reaction is revealed by the occurrence, in the action spectrum of this effect, of peaks at about 560 mμ, 610 mμ, and 640 mμ; a peak at 670 mμ, due to Chl a 670, also is present.  相似文献   

16.
Identification of free radicals was performed for the reaction mixtures of autoxidized 1,2-dilinoleoylphosphatidylcholine (DLPC) with ferrous ions (or DLPC hydroperoxide with ferrous ions) and of DLPC with soybean lipoxygenase using electron spin resonance (ESR), high performance liquid chromatography (HPLC)–ESR and HPLC–ESR–mass spectrometry (MS) combined use of spin trapping technique. ESR measurements of the reaction mixtures showed prominent signals with hyperfine coupling constants (aN=1.58?mT and aHβ=0.26?mT). Outstanding peaks with almost same retention times (autoxidized DLPC, 36.9?min; DLPC hydroperoxide, 35.0?min; DLPC with soybean lipoxygenase, 37.1?min) were observed on the elution profile of the HPLC–ESR analyses of the reaction mixtures. HPLC–ESR–MS analyses of the reaction mixtures gave two ions at m/z 266 and 179, suggesting that 4-POBN/pentyl radical adduct forms in these reaction mixtures.  相似文献   

17.
The experimental hybridizations between fully deoxygenated human and canine hemoglobins and between half-ligated human hemoglobin and canine cyanomethemoglobin show that new two hybrids in addition to the parent hemoglobins were clearly formed in the mixtures at the high concentration of KI. Thus, human deoxyhemoglobin under the present conditions is in an equilibrium with three species, tetramer in equilibrium dimer in equilibrium monomer. This means that the deoxyhemoglobin is in R-T equilibrium, and shifts considerably toward the R state under the present conditions. On the other hand, the half-ligated hemoglobin in 1.5 M KI becomes much more dissociable than the deoxy T state and appears to be completely transformed into the R state. Nevertheless, the co-operativity, n, is still high (n = 2.0).  相似文献   

18.
An improved HPLC-based 32P-postlabeling assay has been developed for the analysis of DNA modified with the food carcinogen 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP). Postlabeled samples are loaded onto a C18 precolumn and adducted bases are retained while excess radioactivity and unmodified DNA bases are eluted directly to waste through a switching valve. The use of this HPLC in-line precolumn purification (HIPP) technique allows entire postlabeled samples to be analyzed without prior removal of inorganic phosphate and unmodified DNA bases. The method has a sample to sample precision of 15% and accuracy of 20%, at adduct levels of 2 adducts/107 bases and shows a linear relationship between signal and adduction levels from 1 adduct per 104 to ≈ 2±1 adducts per 109 bases. Individual postlabeled DNA samples can be analyzed by HPLC in less than 1 h, allowing high throughput. The use of calf-thymus DNA (CT-DNA), highly modified with PhIP, or DNA isolated from mice chronically fed a PhIP-modified diet shows two major PhIP-DNA adduct peaks and three additional minor adduct peaks when labeled under ATP-limiting conditions. Isolation of the HPLC purified peaks and analysis by thin layer chromatography (TLC) matches the five HPLC peaks to the spots typically seen by TLC, including N-(deoxyguanosin-8-yl)-2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (dG-C8-PhIP). Variations in digestion techniques indicate a potential resistance of the PhIP-DNA adducts to the standard enzymatic digestion methods. Attempts at adduct intensification by solid phase extraction, nuclease P1 enrichment or 1-butanol extraction decreased PhIP-DNA adduct peaks and introduced a large early eluting peak. Removal of the 3′-phosphate with nuclease P1 following the kinase labeling reaction simplifies the HPLC profile to one major peak (dG-C8-PhIP monophosphate) with several minor peaks. In addition to the high resolution provided by HPLC separation of the PhIP-DNA adducts, this method can be adjusted for analysis of other DNA adducts and is readily automated for high throughput.  相似文献   

19.
Absorption and fluorescence spectra in the red region of water-soluble chlorophyll proteins, Lepidium CP661, CP663 and Brassica CP673, pigment System II particles of spinach chloroplasts and chlorophyll a in diethylether solution at 25°C were analyzed by the curve-fitting method (French, C.S., Brown, J.S. and Lawrence, M.C. (1972) Plant Physiol. 49, 421–429). It was found that each of the chlorophyll forms of the chlorophyll proteins and the pigment System II particles had a corresponding fluorescence band with the Stokes shift ranging from 0.6 to 4.0 nm.The absorption spectrum of chlorophyll a in diethylether solution was analyzed to one major band with a peak at 660.5 nm and some minor bands, while the fluorescence spectrum was analyzed to one major band with a peak at 664.9 nm and some minor bands. A mirror image was clearly demonstrated between the resolved spectra of absorption and fluorescence. The absorption spectrum of Lepidium CP661 was composed of a chlorophyll b form with a peak at 652.8 nm and two chlorophyll a forms with peaks at 662.6 and 671.9 nm. The fluorescence spectrum was analyzed to five component bands. Three of them with peaks at 654.8, 664.6 and 674.6 nm were attributed to emissions of the three chlorophyll forms with the Stokes shift of 2.0–2.7 nm. The absorption spectrum of Brassica CP673 had a chlorophyll b form with a peak at 653.7 nm and four chlorophyll a forms with peaks at 662.7, 671.3, 676.9 and 684.2 nm. The fluorescence spectrum was resolved into seven component bands. Four of them with peaks at 666.7, 673.1, 677.5 and 686.2 nm corresponded to the four chlorophyll a forms with the Stokes shift of 0.6–4.0 nm. The absorption spectrum of the pigment System II particles had a chlorophyll b form with a peak at 652.4 nm and three chlorophyll a forms with peaks at 662.9, 672.1 and 681.6 nm. The fluorescence spectrum was analyzed to four major component bands with peaks at 674.1, 682.8, 692.0 and 706.7 nm and some minor bands. The former two bands corresponded to the chlorophyll a forms with peaks at 672.1 and 681.6 nm with the Stokes shift of 2.0 and 1.2 nm, respectively.Absorption spectra at 25°C and at ?196°C of the water-soluble chlorophyll proteins were compared by the curve-fitting method. The component bands at ?196°C were blue-shifted by 0.8–4.1 nm and narrower in half widths as compared to those at 25°C.  相似文献   

20.
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   

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