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.
Methyl acetyl phosphate, which was originally synthesized as a site-specific reagent for hydroxybutyrate dehydrogenase [R. Kluger and W.-C. Tsui (1980) J. Org. Chem. 45, 2723], also has an affinity for the binding site for 2,3-diphosphoglycerate in hemoglobin. Three residues in or near this cleft between the beta-chains are acetylated by this reagent, i.e., Val-1, Lys-82, and Lys-144. There is no detectable acetylation of any of the amino groups of the alpha-chain. These results indicate the specificity of methyl acetyl phosphate in its reaction with hemoglobin.  相似文献   

2.
The development of chemical modification agents that reduce the tendency of sickle hemoglobin (HbS) to aggregate represents an important chemotherapeutic goal. Methyl acetylphosphate (MAP) has been reported to bind to the 2,3-diphosphoglycerate (2,3-DPG) binding site of hemoglobin, where it selectively acetylates residues, resulting in increased solubility of HbS. We have prepared [1-(13)C]MAP and evaluated the adduct formation with hemoglobin using (1)H-(13)C HMQC and HSQC NMR studies. These spectra of the acetylated hemoglobin adducts showed 10-11 well resolved adduct peaks, indicating that the acetylation was not highly residue specific. The chemical shift pattern observed is in general similar to that obtained recently using [1'-(13)C]aspirin as the acetylating agent (Xu, A. S. L., Macdonald, J. M., Labotka, R. J., and London, R. E. (1999) Biochim. Biophys. Acta 1432, 333-349). Blocking the 2, 3-DPG binding site with inositol hexaphosphate (IHP) resulted in a selective reduction in intensity of adduct resonances, presumably corresponding to residues located in the 2,3-DPG binding cleft. The pattern of residue protection appeared to be identical to that observed in our previous study using IHP and labeled aspirin. Pre-acetylation of hemoglobin using unlabeled MAP, followed by acetylation with [1'-(13)C]aspirin indicated a general protective effect, with the greatest reduction of intensity for resonances corresponding to acetylated residues in the 2,3-DPG binding site. These studies indicated that both MAP and aspirin exhibit similar, although not identical, acetylation profiles and target primarily the betaLys-82 residue in the 2,3-DPG binding site, as well as sites such as betaLys-59 and alphaLys-90, which are not located in the beta-cleft of hemoglobin.  相似文献   

3.
Reactivity of surface lysyl residues of proteins with a broad range of chemical agents has been proposed to be dependent on the catalytic microenvironment of the residue. We have investigated the acetylation of wild type ubiquitin and of the UbH68N mutant to evaluate the potential contribution of His-68 to the reactivity of Lys-6, which is about 4 A distant. These studies were performed using [1-(13)C]acetyl salicylate or [1,1'-(13)C(2)]acetic anhydride, and the acetylated products were detected by two-dimensional heteronuclear multiple quantum coherence spectroscopy. The results demonstrate that His-68 makes a positive contribution to the rate of acetylation of Lys-6 by labeled aspirin. Additionally, a pair of transient resonances is observed after treatment of wild type ubiquitin with the labeled acetic anhydride but not upon treatment of the H68N mutant. These resonances are assigned to the acetylated His-68 residue. The loss of intensity of the acetylhistidine resonances is accompanied by an increase in intensity of the acetyl-Lys-6 peak, supporting the existence of a transacetylation process between the acetylhistidine 68 and lysine 6 residues located on the protein surface. Hence, this may be the first direct demonstration of a catalytic intermediate forming on the protein surface.  相似文献   

4.
5.
An allosteric modulator of oxygen release in human erythrocytes is 2,3-diphosphoglycerate, but bovine erythrocytes apparently utilize chloride for this purpose since they contain little, if any, 2,3-diphosphoglycerate. In order to identify the sites to which these anions bind, the site-specific acetylating agent, methyl acetyl phosphate, has been employed to compete with these allosteric modulators and to mimic their effects on hemoglobin function. With human hemoglobin A, methyl acetyl phosphate competes with 2,3-diphosphoglycerate and acetylates only Val-1(beta), Lys-82(beta), and Lys-144(beta) within or near the cleft that binds this organic phosphate (Ueno, H., Pospischil, M. A., Manning, J. M., and Kluger, R. (1986) Arch Biochem. Biophys. 244, 795). With bovine hemoglobin, the acetylation is competitive with chloride ion. The sites of acetylation in oxy bovine hemoglobin are Met-1(beta) and Lys-81(beta) and for deoxy bovine hemoglobin, they are Val-1(alpha) and Lys-81(beta). Thus, these sites are expected to be involved in the binding of chloride to bovine hemoglobin. Treatment of either human or bovine hemoglobins with methyl acetyl phosphate under anaerobic conditions leads to a lowering of their oxygen affinity and hence the covalent modifier has the same effect on hemoglobin function as the non-covalent regulators, 2,3-diphosphoglycerate and chloride. The Hill's coefficient of hemoglobin is unaffected by treatment with methyl acetyl phosphate. Under aerobic conditions, specifically acetylated bovine hemoglobin also has a lowered oxygen affinity, and human hemoglobin A shows a slight change in its oxygen affinity. In general, bovine hemoglobin is more responsive than human hemoglobin to both chloride and methyl acetyl phosphate; the latter agent results in a permanent covalent labeling of the protein. Therefore, the results support the idea that methyl acetyl phosphate may be a useful probe for deciphering the sites of binding of anions to proteins.  相似文献   

6.
7.
8.
J D De Caro  M P Chautan  P Rouimi  M Rovery 《Biochimie》1988,70(12):1785-1790
The reactions of lipase (449 amino acid residues) and lipase fragment (336-449) with p-nitrophenyl acetate have been studied from 2 different angles. In previous papers it has been shown that lipase and lipase fragment enzymatically hydrolyze p-nitrophenyl acetate. The amino acid residue of the catalytic site that is temporarily acetylated has not yet been characterized in lipase or lipase fragment. Besides this very fast enzymatic hydrolysis, acetylation reactions may take place on nucleophilic amino acid side-chain groups. In the present report, acetylated amino acid residues whose acetyl linkages were not cleaved after pH 7.5-8.5 incubations have been investigated. Several residues were acetylated in very low proportion, whereas lysine 373 was stoichiometrically acetylated in lipase and in lipase fragment. This specific acetylation may have been favored by the presence of a hydrophobic reversible binding site for p-nitrophenyl acetate near Lys-373. This acetylation did not greatly change the specific activity of lipase towards an emulsion of tributyrylglycerol in the presence of colipase, but under certain conditions it had an effect on the enzymatic hydrolysis of p-nitrophenyl acetate by the lipase fragment.  相似文献   

9.
One of the processes by which CheY, the excitatory response regulator of chemotaxis in Escherichia coli, can be activated to generate clockwise flagellar rotation is by acetyl-CoA synthetase (Acs)-mediated acetylation. Deletion of Acs results in defective chemotaxis, indicating the involvement of Acs-mediated acetylation in chemotaxis. To investigate whether Acs is the sole acetylating agent of CheY, we purified the latter from a delta acs mutant. Mass spectrometry analysis revealed that this protein is partially acetylated in spite of the absence of Acs, suggesting that CheY can be post-translationally acetylated in vivo by additional means. Using [14C]AcCoA in the absence of Acs, we demonstrated that one of these means is autoacetylation, with AcCoA serving as an acetyl donor and with a rate similar to that of Acs-mediated acetylation. Biochemical characterization of autoacetylated CheY and mass spectrometry analysis of its tryptic digests revealed that its acetylated lysine residues are those found in CheY acetylated by Acs, but the acetylation-level distribution among the acetylation sites was different. Like CheY acetylated by Acs, autoacetylated CheY could be deacetylated by Acs. Also similarly to the case of Acs-mediated acetylation, the phosphodonors of CheY, CheA and acetyl phosphate, each inhibited the autoacetylation of CheY, whereas the phosphatase of CheY, CheZ, enhanced it. A reduced AcCoA level interfered with chemotaxis to repellents, suggesting that CheY autoacetylation may be involved in chemotaxis of E. coli. Interestingly, this interference was restricted to repellent addition and was not observed with attractant removal, thus endorsing our earlier suggestion that the signaling pathway triggered by repellent addition is not identical to that triggered by attractant removal.  相似文献   

10.
The current knowledge on biological protein acetylation is confined to acetyl CoA-dependent acetylation of protein catalyzed by specific acetyl transferases and the non-enzymatic acetylation of protein by acetylated xenobiotics such as aspirin. We have discovered a membrane-bound enzyme catalyzing the transfer of acetyl groups from the acetyl donor 7,8-diacetoxy-4-methyl coumarin (DAMC) to glutathione S-transferase 3-3 (GST3-3), termed DAMC:protein transacetylase (TAase). The purified enzyme was incubated with recombinant GST3-3 subunit and DAMC, the modified protein was isolated by sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS–PAGE) in gel digested with trypsin and the tryptic digest was analyzed by mass spectrometry. The N-terminus and six lysines, Lys-51, -82, -124, -181, -191 and -210, were found to be acetylated. The acetylation of GST3-3 described above was not observed in the absence of either DAMC or TAase. These results clearly establish the phenomenon of protein acetylation independent of acetyl CoA catalyzed by a hitherto unknown enzyme (TAase) utilizing a certain xenobiotic acetate (DAMC) as the active acetyl donor.  相似文献   

11.
In this paper we report proton two-dimensional NMR experiments on isolated alpha chains from human hemoglobin A (HbA) in the monocarboxylated state. Several J-correlated and NOE spectra in water or deuterium water and phosphate buffer (100 mM) at 310 K and pH 5.6 were acquired and analysed for the sequential assignment of the proton resonances. In addition, we used the topological data obtained from the crystal structure of alpha subunits in the monocarboxylated HbA tetramer. The assigned resonances correspond to 70% of the amino acid residues. The present results provide information on the tertiary structure of isolated alpha chains in solution, particularly in the heme region. This structure may be compared with that of the a subunits in the tetrameric HbA(CO) in crystal by comparison of observed chemical shifts and those calculated from the X-ray atomic coordinates. Overall, the global folding of the two forms are highly similar. However, this analysis points out several local conformational differences in the heme pocket and the neighboring of the unique Trp residue. Possible explanations of these differences are discussed.  相似文献   

12.
Rat liver nuclei were incubated for various lengths of time in the presence of increasing concentrations of acetyl CoA. The rate of acetylation varied strongly according to the acetyl CoA concentration. A very small part of the histone was acetylated. After incubation in the presence of increasing acetyl CoA concentrations, four apparent Km could be determined. Electrophoretic analysis showed that only histone H3 was acetylated which suggests that each Km corresponds to a sequential acetylation of its lysine residues. This could be correlated with the possible role of histone acetylation in the control of gene activity.  相似文献   

13.
The tyrosine residues of guanidinated horse heart cytochrome c have been specifically acetylated by reaction with N-[1-13C]acetylimidazole (90 atom%). Acetylation was monitored by 13C-NMR spectroscopy. The tyrosine residues were found to show widely varying reactivities ranging from one that is completely and exclusively acetylated at low reagent concentration (residue 67) to one that is acetylated only when the protein is unfolded (residue 97). Homogeneous derivatives were prepared containing one (either residue 67 or 97), three 48, 67 and 74), or four (residues 48, 67, 74 and 97) O-[1-13C]acetyl groups. 13C-NMR spectra of selected derivatives were obtained at pH 5.8, in the presence of cyanide ion, in the ferrous and ferric oxidation states, and after denaturation with 6M guanidine hydrochloride. The O-[1-13C]acetyltyrosyl resonances gave chemical shift values ranging from 171.8 to 176.0 ppm. These resonances were assigned to specific groups based on the known order of reactivity of the tyrosyl side chains toward N-acetylimidazole. The chemical shift of O-[1-13C]acetyltyrosyl 67 was found to be particularly sensitive to changes in protein structure. The proximity of this group to the heme makes it subject to distance-dependent paramagnetic and ring current effects. Acetylation of tyrosyl 74 gives rise to a pH-dependent equilibrium between conformers in the ferric state and a conformation change in the ferrous state. Acetylation of this residue also leads to an absorbance decrease at 695 nm that can be related to the 13C-NMR-detected conformational equilibrium. Addition of cyanide ion abolished this equilibrium.  相似文献   

14.
Jia Z  Cash M  Darvill AG  York WS 《Carbohydrate research》2005,340(11):1818-1825
Eight oligosaccharide subunits, generated by endoglucanase treatment of the plant polysaccharide xyloglucan isolated from the culture filtrate of suspension-cultured tomato (Lycopersicon esculentum) cells, were structurally characterized by NMR spectroscopy. These oligosaccharides, which contain up to three endogenous O-acetyl substituents, consist of a cellotetraose core with alpha-D-Xylp residues at O-6 of the two beta-D-Glcp residues at the non-reducing end of the core. Some of the alpha-D-Xylp residues themselves bear either an alpha-L-Arap or a beta-D-Galp residue at O-2. O-Acetyl substituents are located at O-6 of the unbranched (internal) beta-D-Glcp residue, O-6 of the terminal beta-D-Galp residue, and/or at O-5 of the terminal alpha-L-Arap residue. Structural assignments were facilitated by long-range scalar coupling interactions observed in the high-resolution gCOSY spectra of the oligosaccharides. The presence of five-bond scalar coupling constants in the gCOSY spectra provides a direct method of assigning O-acetylation sites, which may prove generally useful in the analysis of O-acylated glycans. Spectral assignment of these endogenously O-acetylated oligosaccharides makes it possible to deduce correlations between their structural features and the chemical shifts of diagnostic resonances in their NMR spectra.  相似文献   

15.
16.
A "HFPK3" peptide containing the 23 residues of the human immunodeficiency virus (HIV) fusion peptide (HFP) plus three non-native C-terminal lysines was studied in dodecylphosphocholine (DPC) micelles with 2D 1H NMR spectroscopy. The HFP is at the N-terminus of the gp41 fusion protein and plays an important role in fusing viral and target cell membranes which is a critical step in viral infection. Unlike HFP, HFPK3 is monomeric in detergent-free buffered aqueous solution which may be a useful property for functional and structural studies. H alpha chemical shifts indicated that DPC-associated HFPK3 was predominantly helical from I4 to L12. In addition to the highest-intensity crosspeaks used for the first chemical shift assignment (denoted I), there were additional crosspeaks whose intensities were approximately 10% of those used for assignment I. A second assignment (II) for residues G5 to L12 as well as a few other residues was derived from these lower-intensity crosspeaks. Relative to the I shifts, the II shifts were different by 0.01-0.23 ppm with the largest differences observed for HN. Comparison of the shifts of DPC-associated HFPK3 with those of detergent-associated HFP and HFP derivatives provided information about peptide structures and locations in micelles.  相似文献   

17.
M R Busch  J E Mace  N T Ho  C Ho 《Biochemistry》1991,30(7):1865-1877
Assessment of the roles of the carboxyl-terminal beta 146 histidyl residues in the alkaline Bohr effect in human normal adult hemoglobin by high-resolution proton nuclear magnetic resonance spectroscopy requires assignment of the resonances corresponding to these residues. Previous resonance assignments in low ionic strength buffers for the beta 146 histidyl residue in the carbonmonoxy form of hemoglobin have been controversial [see Ho and Russu (1987) Biochemistry 26, 6299-6305; and references therein]. By a careful spectroscopic study of human normal adult hemoglobin, enzymatically prepared des(His146 beta)-hemoglobin, and the mutant hemoglobins Cowtown (beta 146His----Leu) and York (beta 146His----Pro), we have resolved some of these conflicting results. By a close incremental variation of pH over a wide range in chloride-free 0.1 M N-(2-hydroxyethyl)piperazine-N'-2-ethanesulfonic acid buffer, a single resonance has been found to be consistently missing in the proton nuclear magnetic resonance spectra of these hemoglobin variants. The spectra of each of these variants show additional perturbations; therefore, the assignment has been confirmed by an incremental titration of buffer conditions to benchmark conditions, i.e., 0.2 M phosphate, where the assignment of this resonance is unambiguous. The strategy of incremental titration of buffer conditions also allows extension of this resonance assignment to spectra taken in 0.1 M [bis(2-hydroxyethyl)amino]tris(hydroxymethyl)methane buffer. Participation of the beta 146 histidyl residues in the Bohr effect has been calculated from the pK values determined for the assigned resonances in chloride-free 0.1 M N-(2-hydroxyethyl)piperazine-N'-2-ethanesulfonic acid buffer. Our results indicate that the contribution of the beta 146 histidyl residues is 0.52 H+/hemoglobin tetramer at pH 7.6, markedly less than the 0.8 H+/hemoglobin tetramer estimated by study of the mutant hemoglobin Cowtown (beta 146His----Leu) by Shih and Perutz [(1987) J. Mol. Biol. 195, 419-422]. We have found that at least two histidyl residues in the carbonmonoxy form of this mutant have pK values that are perturbed, and we suggest that these pK differences may in part account for this discrepancy. Furthermore, summation of the positive contribution of the beta 146 histidyl residues and the negative contribution of the beta 2 histidyl residues to the maximum Bohr effect measured in 0.1 M N-(2-hydroxyethyl)piperazine-N'-2-ethanesulfonic acid buffer suggests that additional sites in the hemoglobin molecule account for proton release upon ligation greater than the contribution of the beta 146 histidyl residues.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
P Gettins  L W Cunningham 《Biochemistry》1986,25(18):5011-5017
The 1H NMR spectrum of human alpha 2-macroglobulin, Mr 716,000, consists of predominantly extremely broad unresolved resonances but also has nine relatively sharp (delta nu 1/2 less than 25 Hz) resonances from aromatic residues. By treatment of alpha 2-macroglobulin with methylamine, chymotrypsin, and subtilisin, it has been shown that eight of these resonances arise from bait region residues. More specifically, assignment has been made of resonances at 6.80 and 7.11 ppm to the ortho and meta protons, respectively, of tyrosine-685 and tentative assignment of a resonance at 7.29 ppm to the aromatic protons of phenylalanine-684. C2 proton resonances from five histidine residues are also visible. Four of these are attributed to residues in the bait region or immediately adjacent to this, at positions 675, 694, 699, and 704. The sharpness of resonances from bait region residues demonstrates the great flexibility of this region of the polypeptide. It is proposed that the flexible region extends from residue 675 to residue 710. These resonances are all affected by proteolytic cleavage in the bait region but are not influenced by the subsequent conformational rearrangement of the whole protein tetramer. The significance of these findings is discussed in relation to the current structural models of alpha 2-macroglobulin.  相似文献   

19.
Isolated intact rat liver Golgi vesicles utilize [acetyl-3H]coenzyme A to add 3H-O-acetyl esters to sialic acids of internally facing endogenous glycoproteins. During this reaction, [3H]acetate also accumulates in the vesicles, even though the vesicles are impermeant to free acetate. On the other hand, entry of intact AcCoA into the lumen of the vesicles could not be demonstrated, and permeabilization of the vesicles did not alter the reaction substantially (Diaz, S., Higa, H. H., Hayes, B. K., and Varki, A. (1989) J. Biol. Chem. 264, 19416-19426). When vesicles prelabeled with [acetyl-3H] coenzyme A are permeabilized with saponin, we can demonstrate a [3H]acetyl intermediate in the membrane that can transfer label to the 7- and 9-positions of exogenously added free N-acetylneuraminic acid but not to glucuronic acid or CMP-N-acetylneuraminic acid. This labeled acetyl intermediate represents a significant portion of the radioactivity incorporated into the membranes during the initial incubation and cannot be accounted for by nonspecifically "trapped" acetyl-CoA in the permeabilized vesicles. There was no evidence for involvement of acetylcarnitine or acetyl phosphate as an intermediate. The overall acetylation reaction appears to involve two steps. The first step (utilization of exogenous acetyl-CoA to form the acetyl intermediate) is inhibited by coenzyme A-SH (apparent Ki = 24-29 microM), whereas the second (transfer from the acetyl intermediate to sialic acid) is not affected by millimolar concentrations of the nucleotide. Studies with amino acid-modifying reagents indicate that 1 or more histidine residues are involved in the first step of the acetylation reaction. Diethylpyrocarbonate (which can react with both nonsubstituted and singly acetylated histidine residues) also blocks the second reaction, indicating that the acetyl intermediate on both sides of the membrane involves histidine residue(s). Taken together with data presented in the preceding paper, these results indicate that the acetylation of sialic acids in Golgi vesicles may occur by a transmembrane reaction, similar to that described for the acetylation of glucosamine in lysosomes (Bame, K. J., and Rome, L. H. (1985) J. Biol. Chem. 260, 11293-11299). However, several features of this Golgi reaction distinguish it from the lysosomal one, including the nature and kinetics of the reaction and the additional involvement of an essential lysine residue. The accumulation of free acetate in the lumen of the vesicles during the reaction may occur by abortive acetylation (viz. transfer of label from the acetyl intermediate to water). It is not clear if this is an artifact that occurs only in the in vitro reaction.  相似文献   

20.
L W Fung  K L Lin  C Ho 《Biochemistry》1975,14(15):3424-3430
High-resoluiton proton nuclear magnetic resonance spectroscopy at 250 MHz has been used to investigate sickle cell hemoglobin. The hyperfine shifted, the ring-current shifted, and the exchangeable proton resonances suggest that the heme environment and the subunit interfaces of the sickle cell hemoglobin molecule are normal. These results suggest that the low oxygen affinity in sickle cell blood is not due to conformational alterations in the heme environment or the subunit interfaces. The C-2 proton resonances of certain histidyl residues can serve as structural probes for the surface conformation of the hemoglobin molecule. Several sharp resonances in sickle cell hemoglobin are shifted upfield from their positions in normal adult hemoglobin. These upfield shifts, which are observed in both oxy and deoxy forms of the molecule under various experimental conditions, suggest that some of the surface residues of sickle cell hemoglobin are altered and they may be in a more hydrophobic environment as compared with that of normal human adult hemoglobin. These differences in surface conformation are pH and ionic strength specific. In particular, upon the addition of organic phosphates to normal and sickle cell hemoglobin samples, the differences in their aromatic proton resonances diminish. These changes in the surface conformation may, in part, be responsible for the abnormal properties of sickle cell hemoglobin.  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