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.
Previous chemical and structural studies have proposed a major role for Asp-49 in the calcium-mediated activation of phospholipases A2. Recently, a new class of phospholipases A2 has been characterized with a lysine in the place of aspartate at position 49 (Maraganore, J. M., Merutka, G., Cho, W., Welches, W., Kézdy, F. J., and Heinrikson, R. L. (1984) J. Biol. Chem. 259, 13839-13843; Maraganore, J. M., and Heinrikson, R. L. (1986) J. Biol. Chem. 261, 4797-4804). Although both the Lys-49 and Asp-49 phospholipases require calcium for enzymatic activity, the Lys-49 enzymes appear to be unique in their ability to bind phospholipids prior to undergoing calcium-mediated activation. We have successfully crystallized the Lys-49 phospholipase A2 from the venom of the American cottonmouth water moccasin (Agkistrodon piscivorus piscivorus). The crystals are tetragonal, the space group being P4(1)2(1)2 or P4(3)2(1)2 with unit cell dimensions of a = b = 71.05 A, and c = 57.76 A. There is only one molecule in the asymmetric unit and the crystals provide good quality diffraction data to 2.2 A.  相似文献   

2.
A new class of phospholipases A2 that have a lysine at position 49 differ from the more conventional Asp-49 enzymes with respect to the sequential binding of the essential cofactor, calcium, and the substrate, phospholipid, in the formation of the catalytic complex (Maraganore, J.M., Merutka, G., Cho, W., Welches, W., Kézdy, F.J., and Heinrikson, R.L. (1984) J. Biol. Chem. 259, 13839-13843). We report here the complete amino acid sequence of the Lys-49 enzyme from Agkistrodon piscivorus piscivorus. The sequence was determined by automated Edman degradation of the intact, S-carboxymethylcysteinyl protein and of peptides derived therefrom by cleavage with cyanogen bromide, chymotrypsin, trypsin, and endoproteinase Lys-C. Despite several changes at amino acid residues previously considered to be invariant, the Lys-49 enzymes are homologous to the Asp-49 phospholipases. Homology is especially apparent in the following: 1) the pattern of 14 half-cystine residues, 2) conservation of hydrophobic residues which have been shown to encircle the active site, and 3) conservation of Asp-99 and His-48 which have been implicated in the catalytic reaction itself. These observations together with kinetic and binding data imply that the Lys-49 phospholipases have a catalytic mechanism and a three-dimensional architecture similar to those of the Asp-49 enzymes. Modeling of the Lys-49 enzyme based upon the structure of bovine pancreatic phospholipase reveals that the epsilon-amino group of Lys-49 can fit easily in the calcium-binding site and, moreover, that this orientation of a cationic side chain at position 49 could account for the characteristic and novel feature of the Lys-49 phospholipases, i.e. that they are able to form complexes with phospholipid in the absence of calcium.  相似文献   

3.
Summary 1H, 15N and 13C resonance assignments are presented for the group II phospholipase A2 (PLA2) from Agkistrodon piscivorus piscivorus. The secondary structure of the enzyme has been inferred from an analysis of coupling constants, interproton distances, chemical shifts, and kinetics of amide exchange. Overall, the secondary structure of this PLA2 is similar to the crystal structure of the homologous group II human nonpancreatic secretory phospholipase [Scott, D.L., White, S.P., Browning, J.L., Rosa, J.J., Gelb, M.H. and Sigler, P.B. (1991) Science, 254, 1007–1010]. In the group I enzyme from porcine pancreas, the amino-terminal helix becomes fully ordered in the ternary complex of enzyme, lipid micelles and inhibitor. The formation of this helix is thought to be important for the increase in activity of phospholipases on aggregated substrates [Van den Berg, B., Tessari, M., Boelens, R., Dijkman, R., De Haas, G.H., Kaptein, R. and Verheij, H.M. (1995) Nature Struct. Biol., 2, 402–406]. However, the group II enzyme from Agkistrodon piscivorus piscivorus possesses a defined and well-positioned aminoterminal helix in the absence of substrate. Therefore, there is a clear difference between the conformations of group I and group II enzymes in solution. These conformational differences suggest that formation of the amino-terminal helix is a necessary, but not sufficient, step in interfacial activation of phospholipases.Abbreviations PLA2 phospholipase A2 - App-D49 phospholipase from Agkistrodon piscivorus piscivorus - NOE nuclear Overhauser effect  相似文献   

4.
Site-directed mutagenesis studies of bovine pancreatic phospholipase A2 (PLA2, overproduced in Escherichia coli) showed that replacement of surface residue Lys-56 by a neutral or hydrophobic amino acid residue resulted in an unexpected and significant change in the function of the enzyme. The kcat for phosphatidylcholine micelles increases 3-4-fold for K56M, K56I, and K56F and ca. 2-fold for K56N and K56T but does not change for K56R. These results suggest that the side chain of residue 56 has significant influence on the activity of PLA2. In order to probe the structural basis for the enhanced activity, the crystal structures of wild-type and K56M PLA2 were determined by X-ray crystallography to a resolution of 1.8 A. The results suggest that the mutation has not only perturbed the conformation of the side chain of Met-56 locally but also caused conformational changes in the neighboring loop (residues 60-70), resulting in the formation of a hydrophobic pocket by residues Met-56, Tyr-52, and Tyr-69. Docking of a phosphatidylcholine inhibitor analogue into the active site of K56M, according to the structure of the complex of cobra venom PLA2-phosphatidylethanolamine inhibitor analogue [White, S.P., Scott, D. L., Otwinowski, Z., Gleb, M. H., & Sigler, P. (1990) Science 250, 1560-1563], showed that the choline moiety [N(CH3)3]+ is readily accommodated into the newly formed hydrophobic pocket with a high degree of surface complementarity. This suggests a possible interaction between residue 56 and the head group of the phospholipid, explaining the enhanced activities observed when the positively charged Lys-56 is substituted by apolar residues, viz., K56M, K56I, and K56F. Further support for this interpretation comes from the 5-fold enhancement in kcat for the mutant K56E with a negatively charged side chain, where there would be an attractive electrostatic interaction between the side chain of Glu-56 and the positively charged choline moiety. Our results also refute a recent report [Tomasselli, A. G., Hui, J., Fisher, J., Zürcher-Neely, H., Reardon, I.M., Oriaku, E., Kézdy, F.J., & Heinrikson, R.L. (1989) J. Biol. Chem. 264, 10041-10047] that substrate-level acylation of Lys-56 is an obligatory step in the catalysis by PLA2.  相似文献   

5.
Three unique bilin peptides, a beta subunit peptide bearing a doubly linked phycourobilin (PUB), and two gamma subunit peptides with singly linked PUB groups, were obtained by enzymatic degradation of Gastroclonium coulteri R-phycoerythrin. These peptides were shown to have the sequences (Klotz, A. V., and Glazer, A. N. (1985) J. Biol. Chem. 260, 4856-4863): (Formula: see text) The sequence of peptide beta-3T was identical to that previously established for a doubly linked phycoerythrobilin (PEB) peptide derived from a B-phycoerythrin (Lundell, D. J., Glazer, A. N., DeLange, R. J., and Brown, D. M. (1984) J. Biol. Chem. 259, 5472-5480). Secondary ion mass spectrometry of beta-3T yielded a protonated molecular ion of 1629 mass units, the same as that given by the doubly linked PEB peptide (Schoenleber, R. W., Lundell, D. J., Glazer, A. N., and Rapoport, H. (1984) J. Biol. Chem. 259, 5481-5484), indicating that the doubly linked PUB and PEB tetrapyrroles were isomeric structures. High resolution 1H NMR analyses of peptides beta-3T, gamma-BV8, and gamma-DP provided unambiguous structural assignments for the singly and doubly linked PUB chromophores and indicated that the peptides in gamma-BV8 and gamma-DP were linked to ring A. The determination of which peptide fragment is linked to ring A and which to ring D in peptide beta-3T was not achieved in this study. 1H NMR analyses of three PEB-peptides from G. coulteri R-phycoerythrin--alpha-1 Cys(PEB)-Tyr-Arg, alpha-2 Leu-Cys(PEB)-Val-Pro-Arg, and beta-1 Met-Ala-Ala-Cys(PEB)-Leu-Arg--showed that they were identical to previously described corresponding chromopeptides from Porphyridium cruentum B-phycoerythrin, with the peptide linked to ring A of PEB in each instance (Schoenleber, R. W., Lundell, D. J., Glazer, A. N., and Rapoport, H. (1984) J. Biol. Chem. 259, 5485-5489). This is the first documented report on the structure of singly or doubly linked phycourobilins.  相似文献   

6.
The aspartyl residue at position 49 in phospholipases A2 (PLA) has been viewed as a component of the catalytic apparatus because of its involvement in binding the essential cofactor, calcium. We recently discovered a new class of PLA's in which, among other changes in highly invariant residues, Asp-49 is replaced by a lysine (Maraganore et al. (1984) J. Biol. Chem. 259, 13839). These Lys-49 PLA's are also calcium-dependent, but, in contrast to the Asp-49 enzymes, they bind phospholipid strongly in the absence of calcium. Lys-49 PLA's are, therefore, ideal for studying structural and mechanistic aspects of these enzymes. Attempts to modify Lys-49 with the amino group-specific reagent, trinitrobenzenesulfonic acid (TNBS) led to the inactivation of the PLA, but reaction occurred not as expected at position 49, but at Lys-53. These findings lead us to propose a model, applicable to PLA's in general, in which cationic side chains at position 53 in these enzymes participate in phospholipid binding on the path to formation of the catalytic complex. This model serves to explain a number of unresolved observations in the current literature relating to enzyme-substrate interactions in the PLA's.  相似文献   

7.
Glutamine synthetase (Escherichia coli) was incubated with three different reagents that react with lysine residues, viz. pyridoxal phosphate, 5'-p-fluorosulfonylbenzoyladenosine, and thiourea dioxide. The latter reagent reacts with the epsilon-nitrogen of lysine to produce homoarginine as shown by amino acid analysis, nmr, and mass spectral analysis of the products. A variety of differential labeling experiments were conducted with the above three reagents to label specific lysine residues. Thus pyridoxal phosphate was found to modify 2 lysine residues leading to an alteration of catalytic activity. At least 1 lysine residue has been reported previously to be modified by pyridoxal phosphate at the active site of glutamine synthetase (Whitley, E. J., and Ginsburg, A. (1978) J. Biol. Chem. 253, 7017-7025). By varying the pH and buffer, one or both residues could be modified. One of these lysine residues was associated with approximately 81% loss in activity after modification while modification of the second lysine residue led to complete inactivation of the enzyme. This second lysine was found to be the residue which reacted specifically with the ATP affinity label 5'-p-fluorosulfonylbenzoyladenosine. Lys-47 has been previously identified as the residue that reacts with this reagent (Pinkofsky, H. B., Ginsburg, A., Reardon, I., Heinrikson, R. L. (1984) J. Biol. Chem. 259, 9616-9622; Foster, W. B., Griffith, M. J., and Kingdon, H. S. (1981) J. Biol. Chem. 256, 882-886). Thiourea dioxide inactivated glutamine synthetase with total loss of activity and concomitant modification of a single lysine residue. The modified amino acid was identified as homoarginine by amino acid analysis. The lysine residue modified by thiourea dioxide was established by differential labeling experiments to be the same residue associated with the 81% partial loss of activity upon pyridoxal phosphate inactivation. Inactivation with either thiourea dioxide or pyridoxal phosphate did not affect ATP binding but glutamate binding was weakened. The glutamate site was implicated as the site of thiourea dioxide modification based on protection against inactivation by saturating levels of glutamate. Glutamate also protected against pyridoxal phosphate labeling of the lysine consistent with this residue being the common site of reaction with thiourea dioxide and pyridoxal phosphate.  相似文献   

8.
Cystathionine has been reported to undergo enzymatic changes leading to the formation of seven membered cyclic products (Ricci, G., Santoro, L., Achilli, M., Matarese, R. M., Nardini, M., and Cavallini, D. (1983) J. Biol. Chem. 258, 10511-10517; Cavallini, D., Costa, M., Pensa, B., and Coccia, R. (1985) Biochem. Int. 10, 641-646). Gas-chromatographic and mass-spectrometric evidence reported in this paper indicates that the cyclic derivative of cystathionine, 1,4-hexahydrothiazepine-3,5-dicarboxylic acid, here simply named cyclothionine, is a normal component of bovine brain. This finding together with the detection of the same compound in the urine of cystathioninuric patients (Kodama, H., Sasaki, K., Mikasa, H., Cavallini, D., and Ricci, G. (1984) J. Chromatogr. 311, 183-188) supports the conclusion that cystathionine, apart from its role in trans-sulfuration, is converted also into cyclic compounds whose biochemical significance is as yet unknown.  相似文献   

9.
The crystal structure of a lysine 49 variant phospholipase A2 (K49 PLA2) has been determined at 2.0-A resolution. This particular phospholipase A2, purified from the venom of the eastern cottonmouth (Agkistrodon piscivorus piscivorus), a North American pit viper, differs significantly from others studied crystallographically because of replacement of the aspartate residue at position 49, whose side chain is important in calcium binding, by lysine. The crystallographic analysis of K49 PLA2 was undertaken to assess the structural ramifications of this substitution, particularly as they affect the binding mechanism of both the calcium cofactor and the phospholipid substrate. The protein crystals are tetragonal, space group P4(1)2(1)2, with unit cell dimensions of a = b = 71.7 (1) and c = 57.8 (3) A. Preliminary phases were obtained by molecular replacement techniques with a search model derived from the refined 2.5-A structure of a rattle-snake venom phospholipase A2 (Brunie, S., Bolin, J., Gewirth, D., and Sigler, P. B. (1985) J. Biol. Chem. 260, 9742-9749). The starting model gave an initial crystallographic RF of 0.526 (RF = sigma parallel to Fo /-/ Fc parallel to /sigma/Fo/). The structure was refined against all data to 2.0-A resolution. The final RF is 0.158. The final model includes 150 discrete water molecules. The K49 PLA2 model is composed primarily of alpha-helices joined by loops, some of which are quite extensive. Although dissimilarities are observed in the loop regions, the helical portions are very similar to those in other known phospholipase A2 structures. The proposed catalytic center (His48, Tyr73, and Asp99) is also structurally conserved. The region in K49 PLA2 corresponding to the calcium-binding site in other phospholipases A2 is occupied by the epsilon-amino group of lysine 49.  相似文献   

10.
11.
Optical, resonance Raman, and electron paramagnetic resonance spectroscopies have been used to characterize the ligands and spin state of the chloroplast cytochrome b-559. The protein was isolated from both maize and spinach in a low-potential form. The spectroscopic data indicate that the heme iron in both ferric and ferrous cytochrome b-559 is in its low-spin state and ligated in its fifth and sixth coordination positions by histidine nitrogens. Electron paramagnetic resonance data for the purified spinach cytochrome are in good agreement with those determined by Bergstr?m and V?nng?rd [Bergstr?m, J., & V?nng?rd, T. (1982) Biochim. Biophys. Acta 682, 452-456] for a low-potential membrane-bound form of cytochrome b-559. The g values of high-potential cytochrome b-559 are shifted from those of its low-potential forms; this shift is interpreted as arising from a deviation of the planes of the two axial histidine imidazole rings from a parallel orientation. The model is consistent with the physical data and may also account for the facility with which cytochrome b-559 can be converted between low- and high-potential forms. Recent biochemical and molecular biological data [Widger, W. R., Cramer, W. A., Hermodson, M., Meyer, D., & Gullifor, M. (1984) J. Biol. Chem. 259, 3870-3876; Herrmann, R. G., Alt, J., Schiller, D., Cramer, W. A., & Widger, W. R. (1984) FEBS Lett. 179, 239-244] have shown that two polypeptides, one with 83 residues and a second with 39 residues, most likely constitute the protein of the cytochrome.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The porcine pancreatic phospholipase A2-catalyzed hydrolysis of the water-soluble chromogenic substrate 4-nitro-3-octanoyloxybenzoate shows an initial latency phase similar to the one observed in the hydrolysis of aggregated phospholipids by the same enzyme. We report here that during the latency phase the enzyme undergoes a slow, autocatalytic, substrate-level acylation whereby in a few of the catalytic events the scissile octanoyl group of the substrate, normally transferred to water, is transferred to the epsilon-amino group of lysine 56. The N epsilon 56-octanoylphospholipase shows a strong tendency to dimerize in solution and thus may be separated from the monomeric native enzyme by gel filtration. Octanoylation of Lys-56 activates the enzyme some 180-fold toward 4-nitro-3-octanoyloxybenzoate and more than 100-fold toward monolayers of 1,2-didecanoyl-sn-glycero-3-phosphocholine. Acylation also attends the enzymatic hydrolysis of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine with the incorporation of 1 eq of palmitate. Kinetic analysis of the early phase of reaction with 4-nitro-3-octanoyloxybenzoate shows that in this initial step the rate of activation is first order with respect to enzyme and substrate. A much more rapid, autocatalytic activation occurs in the later phases of the reaction where the activation of the enzyme is catalyzed by the activated enzyme itself. These findings with porcine pancreatic phospholipase A2, together with those relative to a snake venom enzyme monomer (Cho, W., Tomasselli, A. G., Heinrikson, R. L., and Kézdy, F. J. (1988) J. Biol. Chem. 263, 11237-11241), strongly support the proposal that interfacial activation of monomeric phospholipases is due to substrate-level autoacylation resulting in fully potentiated dimeric enzymes.  相似文献   

13.
The epsilon-amino group of Lys-166 of Rhodospirillum rubrum ribulosebisphosphate carboxylase/oxygenase was postulated as the essential base which initiates catalysis by abstracting the proton at C-3 of ribulose 1,5-bisphosphate (Hartman, F. C., Soper, T. S., Niyogi, S. K., Mural, R. J., Foote, R. S., Mitra, S., Lee, E. H., Machanoff, R., and Larimer, F. W. (1987) J. Biol. Chem. 262, 3496-3501). To scrutinize this possibility, the site-directed Gly-166 mutant, totally devoid of ribulosebisphosphate carboxylase activity, was examined for its ability to catalyze each of three partial reactions. When carbamylated at Lys-191 (i.e. activated with CO2 and Mg2+), wild-type enzyme catalyzed the hydrolysis of 2-carboxy-3-keto-D-arabinitol 1,5-bisphosphate, the six-carbon reaction intermediate of the carboxylase reaction (Pierce, J., Andrews, T. J., and Lorimer, G. H. (1986a) J. Biol. Chem. 261, 10248-10256). Likewise, when carbamylated at Lys-191, the Gly-166 mutant also catalyzed the hydrolysis of this reaction intermediate. The carbamylated wild type catalyzed the enolization of ribulose 1,5-bisphosphate as indicated by the transfer of 3H radioactivity from [3-3H]ribulose, 1,5-bisphosphate to the medium. However, even when carbamylated at Lys-191, the mutant protein did not catalyze the enolization of ribulose 1,5-bisphosphate. Additionally, unlike the decarbamylated wild-type enzyme, which catalyzed the decarboxylation of 2-carboxy-3-keto-D-arabinitol 1,5-bisphosphate in the absence of Mg2+, the mutant protein was inactive in this partial reaction. These properties exclude the epsilon-amino group of Lys-166 as an obligatory participant in the hydrolysis of 2-carboxy-3-keto-D-arabinitol 1,5-bisphosphate. In contrast, these properties are consistent with the epsilon-amino group of Lys-166 functioning as an acid-base catalyst in the enolization of ribulose 1,5-bisphosphate (when the enzyme is carbamylated) and in the decarboxylation of 2-carboxy-3-keto-D-arabinitol 1,5-bisphosphate (when the enzyme is decarbamylated). Alternatively, Lys-166 may stabilize the transition states of these two partial reactions.  相似文献   

14.
Using modifications of the methods of Bokoch et al. (Bokoch, G.M., Katada, T., Northup, J. K., Ui, M., and Gilman, A. G. (1984) J. Biol. Chem. 259, 3560-3567) and Codina et al. (Codina, J., Hildebrandt, J. D., Sekura, R. D., Birnbaumer, M., Bryan, J., Manclark, C. R., Iyengar, R., and Birnbaumer, L. (1984) J. Biol. Chem. 259, 5871-5886), we have purified a pertussis toxin substrate with the expected characteristics of the inhibitory guanine nucleotide-binding protein (Ni) essentially to homogeneity. The purified protein consists of 3 subunits of Mr 40,000, 35,000, and less than 10,000. The Mr 40,000 band is found, upon close examination, to consist of a poorly resolved doublet. Starting with the membranes from 1,320 g of bovine forebrain we purified the protein some 100-fold with approximately 20% yield to obtain 13 mg of a greater than 95% pure protein. Chromatography on octyl-Sepharose provided efficient separation of Ni from Ns (the stimulatory guanine nucleotide-binding protein). Analytical ultracentrifugation indicates an Mr of 82,000 and a sedimentation coefficient S20,w of 5.1. The protein is able to restore opiate-mediated inhibition of adenylate cyclase to membranes prepared from NG 108-15 cells which had been treated with pertussis toxin. Bovine brain Ni has the enzymatic properties of a low Km GTPase with a turnover number of 0.3 and affinities for nucleotides in the order GppNHp greater than or equal to GTP greater than or equal to GDP much greater than ATP, CTP, UTP, and GMP. Na+ specifically stimulates the GTPase and low concentrations of Mg2+ (less than 50 microM) are inhibitory. Some Mg2+ is apparently necessary because EDTA, but not EGTA, abolishes the GTPase activity.  相似文献   

15.
Previously reported inhibitions of heart lactate dehydrogenase (Guppy, M., and Hochachka, P.W. (1978) J. Biol. Chem. 253, 8465-8469) and muscle pyruvate kinase (Kemp, R.G. (1973) J. Biol. Chem. 248, 3963-3967) by creatine phosphate are due to oxalate which is a contaminant found in some commercial preparations of creatine phosphate.  相似文献   

16.
DNA sequences from the adenovirus 2 genome   总被引:7,自引:0,他引:7  
The sequence of 5,839 nucleotides from the adenovirus 2 genome has been determined and includes the regions between coordinates 32-44% and 66-71%. These regions contain the coding sequences for the 52,55K polypeptide, polypeptide IIIa, penton base, and the N terminus of the 100K polypeptide. Several additional unidentified open reading frames are present, including examples which overlap identified reading frames on the complementary strand and on the same strand. In conjunction with previously published sequences and those described in the accompanying papers (Akusj?rvi, G., Alestr?m, P., Pettersson, M., Lager, M., J?urnvall, H., and Pettersson, U. (1984) J. Biol. Chem. 259, 13976-13979; Alestr?m, P., Akusj?rvi, G., Lager, M., Yeh-kai, L., and Pettersson, U. (1984) J. Biol. Chem. 259, 13980-13985) a complete sequence of 35,937 nucleotide pairs can now be reconstructed for the adenovirus 2 genome.  相似文献   

17.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

18.
Alkaline phosphatase of matrix vesicles isolated from fetal bovine epiphyseal cartilage was purified to apparent homogeneity using monoclonal antibody affinity chromatography. The enzyme from the butanol extract of matrix vesicles bound specifically to the immobilized antibody-Sepharose in the presence of 2% Tween 20 whereas the major portion of nonspecific protein was removed by this single step. Of various agents tested, 0.6 M 2-amino-2-methyl-1-propanol, pH 10.2, was the most effective in eluting 80-100% of the enzyme initially applied. Both Tween 20 and 2-amino-2-methyl-1-propanol associated with the eluted enzyme were effectively removed by the sequential application of DEAE-cellulose and Sepharose CL-6B chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the enzyme preparation treated with sodium dodecyl sulfate and mercaptoethanol showed the presence of a dominant band (using silver staining) corresponding to a molecular weight of 81,000. This molecular weight was nearer reported values for rat liver (Ohkubo, A., Langerman, N., and Kaplan, M. M. (1974) J. Biol Chem. 249, 7174-7180) and porcine kidney (Cathala, G., Brunel, C., Chapplet-Tordo, D., and Lazdunski, M. (1975) J. Biol. Chem. 250, 6040-6045) alkaline phosphatase, than to previously reported values for chicken (Cyboron, G. W., and Wuthier, R. E. (1981) J. Biol. Chem. 256, 7262-7268) and fetal calf (Fortuna, R., Anderson, H. C., Carty, R. P., and Sajdera, S. W. (1980) Calcif. Tissue Int. 30, 217-225) cartilage matrix vesicle alkaline phosphatase. The purified alkaline phosphatase was activated by micromolar Mg2+. The amino acid composition of cartilage alkaline phosphatase was found to be similar to that previously described for porcine kidney (Wachsmuth, E. D., and Hiwada, K. (1974) Biochem. J. 141, 273-282). Double immunoprecipitation data indicated that monoclonal antibody against cartilage alkaline phosphatase cross-reacted with fetal bovine liver or kidney enzyme but failed to react with calf intestinal or rat cartilage enzyme. Thus these observations suggest that alkaline phosphatase of matrix vesicles from calcifying epiphyseal cartilage is a liver-kidney-bone isozyme.  相似文献   

19.
The complex between ferredoxin-NADP+ oxidoreductase and its proposed membrane-binding protein (Vallejos, R. H., Ceccarelli, E., and Chan, R. (1984) J. Biol. Chem. 259, 8048-8051) was isolated from spinach thylakoids and compared with isolated cytochrome b/f complex containing associated ferredoxin NADP+ oxidoreductase (Clark, R. D., and Hind, G. (1983) J. Biol. Chem. 258, 10348-10354). There was no immunological cross-reactivity between the 17.5-kDa binding protein and an antiserum raised against the 17-kDa polypeptide of the cytochrome complex. Association of ferredoxin-NADP+ oxidoreductase with the binding protein or with the thylakoid membrane gave an allotopic shift in the pH profile of diaphorase activity, as compared to the free enzyme. This effect was not seen in enzyme associated with the cytochrome b/f complex. Identification of the 17.5-kDa binding protein as the 17-kDa component of the cytochrome b/f complex is ruled out by these results.  相似文献   

20.
Phospholipases A(2) (PLA(2)) are potent regulators of the inflammatory response. We have observed that Group IV cPLA(2) activity is required for the production of superoxide anion (O(2)(-)) in human monocytes [Li Q., Cathcart M.K. J. Biol. Chem. 272 (4) (1997) 2404-2411.]. We have previously identified PKCalpha as a kinase pathway required for monocyte O(2)(-) production [Li Q., Cathcart M.K. J. Biol. Chem. 269 (26) (1994) 17508-17515.]. We therefore investigated the potential interaction between PKCalpha and cPLA(2) by evaluating the requirement for specific PKC isoenzymes in the process of activating cPLA(2) enzymatic activity and protein phosphorylation upon monocyte activation. We first showed that general PKC inhibitors and antisense oligodeoxyribonucleotides (ODN) to the cPKC group of PKC enzymes inhibited cPLA(2) activity. To distinguish between PKCalpha and PKCbeta isoenzymes in regulating cPLA(2) protein phosphorylation and enzymatic activity, we employed our previously characterized PKCalpha or PKCbeta isoenzyme-specific antisense ODN [Li Q., Subbulakshmi V., Fields A.P., Murray, N.R., Cathcart M.K., J. Biol. Chem. 274 (6) (1999) 3764-3771]. Suppression of PKCalpha expression, but not PKCbeta expression, inhibited cPLA(2) protein phosphorylation and enzymatic activity. Additional studies ruled out a contribution by Erk1/2 to cPLA(2) phosphorylation and activation. We also found that cPLA(2) co-immunoprecipitated with PKCalpha and vice versa. In vitro studies demonstrated that PKCalpha could directly phosphorylate cPLA(2).and enhance enzymatic activity. Finally, we showed that addition of arachidonic acid restored the production of O(2)(-) in monocytes defective in either PKCalpha or cPLA(2) expression. Taken together, our data suggest that PKCalpha, but not PKCbeta, is the predominant cPKC isoenzyme required for cPLA(2) protein phosphorylation and maximal induction of cPLA(2) enzymatic activity upon activation of human monocytes. Our data also support the concept that the requirements for PKCalpha and cPLA(2) in O(2)(-) generation are solely due to their seminal role in generating arachidonic acid.  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