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.
We have synthesized and tested a new series of compounds that inhibit the polymerization of deoxyhemoglobin S by noncovalent interaction. They consist of three structural elements: A p-aminobenzoyl residue to anchor the compound in the central cavity between the beta chains, a number of glutamates in gamma linkage to provide tight binding, and one or two hydrophobic amino acid residues which block the intermolecular hydrophobic interaction of valine beta 6. The most active compound was p-aminobenzoyl-(gamma-Glu)5-Phe-Phe. It increases the solubility of deoxy-HbS by a factor of 1.3 at a concentration of only 5-6 mM and is effective even in the presence of physiological concentrations of 2,3-diphosphoglycerate.  相似文献   

2.
The adult greater Kudu antelope has two hemoglobin components, Hb A and Hb B, with one alpha and two beta chains. The complete amino-acid sequences of these three chains are presented. The two beta chains differ only in one residue at position 16 (Gly----Ser) and may be the product of two allelic genes. The primary structure of the chains was determined by sequencing the tryptic peptides after their isolation from the tryptic digest of the chains by high performance liquid chromatography. The alignment of these peptides was deduced from homology with the chains of bovine hemoglobin. Between the Kudu hemoglobins and those of cattle a high degree of homology was found.  相似文献   

3.
I M Russu  A K Lin  C P Yang  C Ho 《Biochemistry》1986,25(4):808-815
High-resolution proton nuclear magnetic resonance spectroscopy and relaxation techniques have been used to investigate the interactions of sickle cell hemoglobin (Hb S) and human normal adult hemoglobin (Hb A) with p-bromobenzyl alcohol, L-phenylalanine, L-tryptophan, and L-valine. With the exception of valine, all these compounds inhibit the polymerization of deoxy-Hb S [Noguchi, C. T., & Schechter, A. N. (1978) Biochemistry 17, 5455)). Using transferred nuclear Overhauser effects among the proton resonances of the compound of interest and the corresponding longitudinal relaxation rates (T1(-1], we have shown that the binding of each of the compounds investigated to deoxy-Hb S is comparable to that to deoxy-Hb A. Intermolecular transferred nuclear Overhauser effects have been observed between proton resonances of the anti-sickling compounds and specific protons situated in the heme pockets of Hb. On the basis of these results, we suggest that one binding site, common to all compounds with anti-sickling activity, is at or near the heme pockets in the alpha and beta chains of both deoxy-HB S and deoxy-Hb A. The proton T1(-1) values of the histidyl residues situated over the surface of the hemoglobin molecule indicate that a second binding site is located at or near the beta 6 position, containing the mutation in Hb S (beta 6Glu----Val). The binding of the compounds investigated to the latter site induces conformational changes in the amino-terminal domains of the beta chains.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The adult Grand Galago (Galago crassicaudatus) was found to have two hemoglobin components (Hb I and Hb II) which were separated by carboxymethyl cellulose column chromatography. The alpha and beta chains of each component were isolated. The tryptic peptides of the alpha and beta chains were each isolated and sequenced by the conventional method. The alignment of these peptides in each chain was deduced from the homology of their sequences with that of human adult hemoglobin. The alpha chains from Hb I and Hb II were considered to be identical. On the other hand, there was only one amino-acid difference between the two beta chains at the 125th residue from the N-terminus.  相似文献   

5.
The adult Stump-Tail Monkey (Macaca speciosa) was found to have two major hemoglobin components (Hb 1 and Hb 2) which were separated by carboxymethyl cellulose column chromatography. The tryptic peptides of the alpha and beta chains from the two components were isolated and sequenced. The peptides were aligned based on the homology of their sequences with that of human adult hemoglobin. Only one amino-acid difference was found between the alpha chains from Hb 1 and Hb 2 at the 15th position from the N-terminus. On the other hand, the beta chains from the two hemoglobin components were considered to be identical.  相似文献   

6.
Peptide-membrane interactions are important for understanding the binding, partitioning, and folding of membrane proteins; the activity of antimicrobial and fusion peptides; and a number of other processes. We describe molecular dynamics simulations (10-25 ns) of two pentapeptides Ace-WLXLL (with X = Arg or Lys side chain) (White, S. H., and Wimley, W.C. (1996) Nat. Struct. Biol. 3, 842-848) in water and three different membrane mimetic systems: (i) a water/cyclohexane interface, (ii) water-saturated octanol, and (iii) a solvated dioleoylphosphatidylcholine bilayer. A salt bridge is found between the protonated Arg or Lys side chains with the carboxyl terminus at the three interfaces. In water/cyclohexane, the salt bridge is most exposed to the water phase and least stable. In water/octanol and the lipid bilayer systems, the salt bridge once formed persists throughout the simulations. In the lipid bilayer, the salt bridge is more stable when the peptide penetrates deeper into the bilayer. In one of two peptides, a cation-pi interaction between the Arg and the Trp side chains is stable in the lipid bilayer for about 15 ns before breaking. In all cases, the conformations of the peptides are restricted by their presence at the interface and can be assigned to a few major conformational clusters. Side chains facing the water phase are most mobile. In the lipid bilayer, the peptides remain in the interface area, where they overlap with the carbonyl area of the lipid bilayer and perturb the local density profile of the bilayer. The tryptophan side chain remains in the water-lipid interface, where it interacts with the lipid choline group and forms hydrogen bonds with the ester carbonyl of the lipid and with water in the interface.  相似文献   

7.
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   

8.
A low P50 value in a fresh red blood cell suspension was discovered in a polycythemic patient (Hb 19 g X dl-1). Routine acid and alkaline electrophoreses of the hemolysate were identical to normal hemolysate. Isoelectrofocusing (pH gradient 6-8) did not reveal any abnormal band whether performed with the fully liganded or deoxygenated samples. Precise analyses of the oxygen dissociation curves of the propositus' red cells demonstrated a biphasic Hill plot, a normal Bohr effect and low interaction with 2,3-bisphosphoglycerate (2,3-DPG). Studies on the unfractionated hemolysate confirmed these observations and the inhibition of the effect of organic phosphates. Structural studies were carried out on the mixture of beta A + beta X chains and revealed the presence of two beta Tp14 peptides. Sequencing the abnormal beta Tp14 peptide showed the substitution Ala----Thr of the beta 140 (H18) residue. This new variant was named Hb Saint-Jacques. Examination of the three dimensional model of HbAo indicates that the substitution beta 140 (H18) Ala----Thr induces van der Waals interactions with the nearby lysine-82 (EF6) and leucine-81 (EF5) and a displacement of the EF corner of the beta chains. This is likely to change the normal position of the lysine-82 (EF6), a major anionic binding site in the central cavity between the two beta chains. Functional studies confirm the interpretation of a steric hindrance inhibiting the binding of large organic phosphates to Hb Saint-Jacques.  相似文献   

9.
I M Russu  S S Wu  K A Bupp  N T Ho  C Ho 《Biochemistry》1990,29(15):3785-3792
High-resolution 1H and 31P nuclear magnetic resonance spectroscopy has been used to investigate the binding of 2,3-diphosphoglycerate to human normal adult hemoglobin and the molecular interactions involved in the allosteric effect of the 2,3-diphosphoglycerate molecule on hemoglobin. Individual hydrogen ion NMR titration curves have been obtained for 22-26 histidyl residues of hemoglobin and for each phosphate group of 2,3-diphosphoglycerate with hemoglobin in both the deoxy and carbonmonoxy forms. The results indicate that 2,3-diphosphoglycerate binds to deoxyhemoglobin at the central cavity between the two beta chains and the binding involves the beta 2-histidyl residues. Moreover, the results suggest that the binding site of 2,3-diphosphoglycerate to carbonmonoxyhemoglobin contains the same (or at least some of the same) amino acid residues responsible for binding in the deoxy form. As a result of the specific interactions with 2,3-diphosphoglycerate, the beta 2-histidyl residues make a significant contribution to the alkaline Bohr effect under these experimental conditions (up to 0.5 proton/Hb tetramer). 2,3-Diphosphoglycerate also affects the individual hydrogen ion equilibria of several histidyl residues located away from the binding site on the surface of the hemoglobin molecule, and, possibly, in the heme pockets. These results give the first experimental demonstration that long-range electrostatic and/or conformational effects of the binding could play an important role in the allosteric effect of 2,3-diphosphoglycerate on hemoglobin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The thermodynamic binding parameters and crystal structure for streptavidin-peptide complexes where the peptide sequences were obtained by random screening methods are reported. The affinities between streptavidin and two heptapeptides were determined by titrating calorimetric methods [Phe-Ser-His-Pro-Gln-Asn-Thr, Ka = 7944 (+/- 224) M-1, delta G degrees = -5.32 (+/- 0.01) kcal/mol, and delta H degrees = -19.34 (+/- 0.48) kcal/mol; His-Asp-His-Pro-Gln-Asn-Leu, Ka = 3542 (+/- 146) M-1, delta G degrees = -4.84 (+/- 0.03) kcal/mol, and delta H degrees = -19.00 (+/- 0.64) kcal/mol]. The crystal structure of streptavidin complexed with one of these peptides has been determined at 2.0-A resolution. The peptide (Phe-Ser-His-Pro-Gln-Asn-Thr) binds in a turn conformation with the histidine, proline, and glutamine side chains oriented inward at the biotin-binding site. A water molecule is immobilized between the histidine and glutamine side chains of the peptide and an aspartic acid side chain of the protein. Although some of the residues that participate in binding biotin also interact with the screened peptide, the peptide adopts an alternate method of utilizing binding determinants in the biotin-binding site of streptavidin.  相似文献   

11.
One-electron oxidation of six different c-type lysozymes from hen egg white, turkey egg white, human milk, horse milk, camel stomach and tortoise was studied by gamma- and pulse-radiolysis. In the first step, one tryptophan side chain is oxidized to indolyl free radical, which is produced quantitatively. As shown already, the indolyl radical subsequently oxidizes a tyrosine side chain to the phenoxy radical in an intramolecular reaction. However this reaction is not total and its stoichiometry depends on the protein. Rate constants also vary between proteins, from 120 x s(-1) to 1000 x s(-1) at pH 7.0 and room temperature [extremes are hen and turkey egg white (120 x s(-1)) and human milk (1000 x s(-1))]. In hen and turkey egg white lysozymes we show that another reactive site is the Asn103-Gly104 peptidic bond, which gets broken radiolytically. Tryptic digestion followed by HPLC separation and identification of the peptides was performed for nonirradiated and irradiated hen lysozyme. Fluorescence spectra of the peptides indicate that Trp108 and/or 111 remain oxidized and that Tyr20 and 53 give bityrosine. Tyr23 appears not to be involved in the process. Thus new features of long-range intramolecular electron transfer in proteins appear: it is only partial and other groups are involved which are silent in pulse radiolysis.  相似文献   

12.
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   

13.
Low-density lipoproteins floating between densities 1-006 and 1-063 g cm-3 were isolated by centrifugation of blood serum obtained from 24-h fasted pigs (Sus domesticus). This lipoprotein fraction contained two components with Sf 1-063 values of 3-4 and 2-3 at 20 degrees C when examined by analytical ultracentrifugation. Delipidation of the lipoprotein yielded 15% recovery of soluble protein. Chromatography on Sephadex G100 in 8 M urea of these delipidation products yielded three fractions of different sizes which were present in both native and succinylated apoproteins. These fractions from the succinylated apolipoproteins were further characterized. A polypeptide fraction comprising 70% of the total protein had an apparent molecular weight of 34000 and contained greater amounts of amino acids with hydrophobic side chains than did the second fraction of apparent molecular weight 22000 which contained 15% of the protein. The third fraction of apparent molecular weight 12500 contained 15% of the protein.  相似文献   

14.
The pulse-radiolysis technique has been introduced because it permits a rapid reduction (in a few microseconds) of one heme group of the methemoglobin tetramer by hydrated electrons. The kinetics of the binding of oxygen to this particular valence intermediate (Hb3+) with one reduced alpha or beta subunit has been studied. It appears that the hydrated electrons preferentially reduce one type of subunit of methemoglobin at acid and neutral pH-values as is shown by the biphasic behaviour of Hb3+ on oxygenation. The second-order on-rate constants measured for the binding of oxygen to Hb3+ are 14 +/- 3 mM-1 ms-1 and 56 +/- 9 mM-1 ms-1, respectively. The relative contribution of the faster fraction is about 0.63 +/- 0.08 of the total oxygenation process. A comparison of the kinetic absorbance difference spectrum for the reduction of methemoglobin with the static difference spectrum of deoxyhemoglobin and methemoglobin in the Soret-region revealed a decreased absorbance of the unliganded subunit of Hb3+ at 430 nm. This fact suggests that Hb3+ is in the relaxed quaternary conformation, which is in agreement with the observed on-rate constants.  相似文献   

15.
The oxygen binding properties of human hemoglobin are appreciably altered by the nicotinamide dinucleotides NADH, NADP+, and NADPH. These cofactors are important in the control of many metabolic pathways and in providing reductive potential for a number of enzymatic reactions, including in vivo reduction of methemoglobin. Specific binding of these cofactors to hemoglobin and their potential for acting as allosteric modifiers of hemoglobin function have not been previously recognized. Detailed oxygen binding studies utilizing a thin-layer method suggest that the nicotinamide dinucleotides bind with high affinity to the deoxyhemoglobin tetramer at the beta chain anion-binding site and stabilize the low affinity "T-state" conformation. Stripped Hb A in 0.05 M N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid (HEPES) buffer, pH 6.5, at 20 degrees C is half-saturated at a pO2 of 1.6 mm Hg. In the presence of 0.5 mM NADH, NADP+, or NADPH, the P50 is raised to 3.8, 7.1, and 12.5 mm Hg, respectively. The Bohr factor for stripped Hb A in 0.05 M HEPES buffer is sensitive to these effectors and is raised from 0.25 to about 0.65 by the addition of NADPH. The data suggest that routine use of these effectors in studies of human hemoglobin variants or the allosteric mechanism of Hb A be considered carefully. The relatively low intraerythrocytic levels of the nicotinamide dinucleotides in relation to hemoglobin dictate that these cofactors cannot significantly affect in vivo oxygen delivery. However, the converse is theoretically possible. The binding of the cofactors to hemoglobin and the preferential binding of their reduced forms may affect cofactor-dependent metabolic processes in red blood cells.  相似文献   

16.
J A Barden  B E Kemp 《Biochemistry》1987,26(5):1471-1478
The amino acid residues 114-118 in actin were found to be implicated strongly in the binding of nucleotide, and as would be expected for such an important binding site, they are located in a completely conserved region of the actin sequence. A 19-residue peptide with the actin sequence 106-124 was synthesized in order to span the putative triphosphate binding site. Proton NMR spectra of the actin peptide 114-118 in the presence and absence of ATP indicated that Arg-116 and Lys-118 are particularly involved in binding ATP. A strong binding of ATP to the peptide 106-124 also was measured. Tripolyphosphate bound to the peptide 106-124 somewhat more weakly than ATP. Binding involved residues 115-118 and 121-124, indicating the presence of a reverse turn between these segments. Proton resonances were assigned by using two-dimensional double quantum correlated spectroscopy, one-dimensional spin decoupling techniques, one-dimensional nuclear Overhauser enhancement difference spectroscopy, and pH titration. The alpha CH resonances of Ala-3 and Asn-6 are markedly shifted downfield with respect to values in small unstructured peptides due to their close proximity to the side chains of Pro-4 and Pro-7, respectively. Several other resonances display chemical shifts which are indicative of a structured environment. Assignment of the amide proton resonances in H2O and measurements of the coupling constant 3JHNCH and the chemical shifts of the amide protons reveal that much of the synthetic peptide, particularly the backbone, exhibits a highly structured environment and represents a good model for the triphosphate binding site in actin.  相似文献   

17.
The Escherichia coli periplasmic dipeptide binding protein functions in both peptide transport and taxis toward peptides. The structure of the dipeptide binding protein in complex with Gly-Leu (glycyl-L-leucine) has been determined at 3.2 A resolution. The binding site for dipeptides is designed to recognize the ligand's backbone while providing space to accommodate a variety of side chains. Some repositioning of protein side chains lining the binding site must occur when the dipeptide's second residue is larger than leucine. The protein's fold is very similar to that of the Salmonella typhimurium oligopeptide binding protein, and a comparison of the structures reveals the structural basis for the dipeptide binding protein's preference for shorter peptides.  相似文献   

18.
The delta selectivity of the opioid heptapeptides deltorphin I and II has been attributed to the C-terminal 'address' domain, the hydrophobic Val(5)-Val(6) residues apparently playing a topographical role. We now report the synthesis, opioid binding affinities, and a QSAR study of a series of peptides in which one of the valine side chains was altered. QSAR analyses included previously published models for a binding pocket interaction and an optimum size (Schullery, S.; Mohammedshah, T.; Makhlouf, H.; Marks, E.; Wilenkin, B.; Escobar, S.; Mousigian, C.; Heyl, D. Bioorg. Med. Chem. 1997, 5, 2221), and a new approach for backbone conformational effects using Langevin dynamics simulation (PM3 semi-empirical force field) of an isolated peptide fragment containing the side chain and flanking peptide bonds. No evidence is found of binding pocket interactions or optimum size for either the position-5 or -6 side chain. Rather, delta binding is generally disfavored while mu binding is either unaffected (position-5) or favored (position-6) by larger side chains. The dynamics results provide evidence of similar 'local' conformation roles for the positions 5 and 6 side chains. Specifically, delta binding is favored by side chains that maximize the extension of the backbone, measured as the through-space distance between peptide fragment ends, the angle between lines connecting the alpha-carbon with fragment ends, or the difference between the psi and phi peptide angles.  相似文献   

19.
Summary. Numerous indolyl amino acids and their derivatives inhibited arginase activity. The inhibition was found to be non-competitive, – at least partly – allosteric, and independent on manganese ions in the active site, and it cannot be explained by the dissociation of arginase homotrimers. Indole alone is weakly inhibitory; however, the presence of three-carbon side chains and their net charges is favorable for the inhibition. The binding of the inhibitory compounds caused only minor changes in the steric structure of arginase: a slight increase in α-helix content was detected by circular dichroism together with a decrease in parallel pleated sheet and β-turn sections. A slight alteration in the tertiary structure was also found using tryptophane fluorescence studies, but buried apolar side chains were not transposed to the protein surface. Computer studies that were performed did not provide additional structural information. Authors’ address: András Hrabák, Department of Medical Chemistry, Molecular Biology and Patho-biochemistry, Semmelweis University Medical School, Budapest, VIII. Puskin u. 9., H-1444 POB 260, Hungary  相似文献   

20.
The saccharide binding ability of the low affinity (LA-) binding site of ricin D was abrogated by N-bromosuccinimide (NBS)-oxidation, while in the presence of lactose the number of tryptophan residues eventually oxidized decreased by 1 mol/mol and the saccharide binding ability was retained (Hatakeyama et al., (1986) J. Biochem. 99, 1049-1056). Based on these findings, the tryptophan residue located at the LA-binding site of ricin D was identified. Two derivatives of ricin D which were modified with NBS in the presence and absence of lactose were separated into their constituent polypeptide chains (A- and B-chains), respectively. The modified tryptophan residue or residues was/were found to be contained in the B-chain, but not in the A-chain. From lysylendopeptidase and chymotryptic digests, peptides containing oxidized tryptophan residues were isolated by gel filtration on Bio-Gel P-30 and HPLC. Analysis of the peptides containing oxidized tryptophan revealed that three tryptophan residues at positions 37, 93, and 160 on the B-chain were oxidized in the inactive derivative of ricin D, in which the saccharide binding ability of the LA-binding site was abrogated by NBS-oxidation. On the other hand, the modified residues were determined to be tryptophans at positions 93 and 160 in the active derivative of ricin D which was modified in the presence of lactose, indicating that upon binding with lactose, the tryptophan residue at position 37 of the B-chain was protected from NBS-oxidation. From these results, it is suggested that tryptophan at position 37 on the B-chain is the essential residue for saccharide binding at the LA-binding site of ricin D.  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