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.
Effects of Ca2+ on the kinetic parameters for the hydrolysis of mixed micelles of 1,2-dipalmitoyl-sn-glycero-3-phosphorylcholine (diC16PC) with Triton X-100, catalyzed by a cobra (Naja naja atra) (Group I) and a Habu (Trimeresurus flavoviridis) (Group II) PLA2s, were studied and compared with the results reported for other Group I and II enzymes. The substrate bindings to Group I enzymes were independent of the Ca2+ binding, whereas the substrate bindings to Group II enzymes were facilitated more than 10 times by the Ca2+ binding to the enzymes. The result for Group II enzymes, but not Group I enzymes, seemed compatible with the hypothesis for interpreting the catalytic mechanism that an intermediate complex should be stabilized by the coordination of the bound Ca2+ with the phosphoryl group and the carbonyl oxygen atom of the ester bond at the sn-2 position of the bound substrate molecule [Verheij et al. (1980) Biochemistry 19, 743-750 and (1981) Rev. Physiol. Biochem. Pharmacol. 91, 91-203]. The pH dependence of the kinetic parameters for the hydrolysis of the mixed micellar diC16PC, catalyzed by the cobra (N. naja atra) (Group I) and Habu (T. flavoviridis) (Group II) PLA2s, was also studied. The pK values of the catalytic group, His 48, and Tyr 52 for N. naja atra PLA2, shifted from 7.25 to 7.70 and from 10.30 to 10.85, respectively, and the corresponding values for T. flavoviridis PLA2 shifted from 5.80 to 6.95 and from 10.10 to 10.76, respectively, on binding of the micellar substrates to the enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The hydrolysis of 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by a cobra (Naja naja atra) venom phospholipase A2, was studied at 25 degrees C ionic strength 0.1 in the presence of 3-10 mM Ca2+, which can saturate the Ca2+-binding site of the enzyme. The initial velocity data, obtained at various concentrations of the substrate below the critical micellar concentration (cmc), were analyzed according to the Michaelis-Menten equation. The Km value was practically independent of pH (between pH 6.75 and 10.30). This finding was consistent with the result of a direct binding study on monodispersed n-alkylphosphorylcholines (Teshima et al. (1981) J. Biochem. 89, 1163-1174). The hydrolysis of the substrate was competitively inhibited by the presence of monodispersed n-dodecylphosphorylcholine (n-C12PC). These results indicated that the substrate and n-C12PC compete for the same site on the enzyme molecule. The pH dependence curve of the kinetic parameter, kcat/Km, exhibited three transitions, below pH 8, between pH 8 and 9.5, and above pH 10. The analysis indicated the participation of three ionizable groups with pK values of 7.25, 8.50, and 10.4. The deprotonation of the first group and the protonation of the third group were found to be essential for the catalysis. The first group was assigned as His 48 in the active site on the basis of its pK value, which had been determined from the pH dependence of the binding constant of Ca2+ (Teshima et al. (1981) J. Biochem. 89, 13-20).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The phosphatidylcholine-hydrolyzing phospholipase C, so-called "phospholipase C" (PLC), was isolated from the culture of Bacillus cereus strain IAM 1208. The amino-acid composition and partial N-terminal sequence of the purified enzyme were in good agreement with those expected from the nucleotide sequence for a PLC of strain ATCC 10987 [Johansen et al. (1988) Gene 65, 293-304]. The chain-length dependence of kinetic parameters for the PLC-catalyzed hydrolysis of monodispersed short-chain phosphatidylcholines (diCNPC, N = 3-6) was studied by a pH-stat assay method at 25 degrees C, pH 8.0, and ionic strength 0.2 in the presence of saturating amounts of Zn2+ (0.1 mM). The result was compared with those for snake venom phospholipases A2 [Teshima et al. (1989) J. Biochem. 106, 518-527]. It was found that the interaction of the PLC with the head group of the substrate molecule is very important for the binding. The pH dependences of kinetic parameters for the hydrolysis of monodispersed diC5PC and mixed micelles of diC16PC with Triton X-100 were also studied under the same conditions. An ionizable group, whose pK value is perturbed from 7.77 to 8.30 by substrate binding, was found to be essential to the catalysis. This group was tentatively assigned to His 14 on the basis of the results on X-ray crystallographic and chemical modification studies [Hough et al. (1989) Nature 338, 357-360 and Little (1977) Biochem. J. 167, 399-404].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The pH dependence of the chemical reaction rate of p-bromophenacyl bromide (BPB) with His 48 of cobra (Naja naja atra) venom phospholipase A2, in which the alpha-NH2 group had been selectively modified to an alpha-keto group, was studied at 25 degrees C and ionic strength 0.1 in the absence of Ca2+. The pH-dependence curve was monophasic with a midpoint at pH 7.9, which corresponds to the pK value of His 48 of the alpha-NH2-modified enzyme, whereas the curve for the intact enzyme was biphasic, indicating participation of two ionizable groups with pK values of 7.3 and 8.55 (Teshima et al. (1982) J. Biochem. 91, 1778-1788). These two groups were thus identified as His 48 and the alpha-NH2 group, respectively. The pH dependence of the binding constant of Ca2+ to the alpha-NH2-modified enzyme was studied at 25 degrees C and ionic strength 0.1 by measuring the tryptophyl fluorescence changes. The pH-dependence curve was very similar to that for the intact enzyme (Teshima et al. (1981) J. Biochem. 89, 13-20), and it was interpreted in terms of participation of His 48 and Asp 49 (pK 5.4). The absence of participation of the alpha-NH2 group in the Ca2+ binding was thus confirmed. Bindings of monodispersed n-dodecylphosphorylcholine (n-C12PC) and micellar n-hexadecylphosphorylcholine (n-C16PC) to the alpha-NH2-modified enzyme were studied at 25 degrees C and ionic strength 0.1 by the aromatic circular dichroism (CD) and tryptophyl fluorescence methods, respectively. The binding constant of the monodispersed substrate was very similar to that for the intact enzyme (Teshima et al. (1981) J. Biochem. 89, 1163-1174). The binding constant of the micellar substrate to the modified enzyme in the presence of Ca2+ was also very similar to that for the intact enzyme-Ca2+ complex (Teshima et al. (1983) J. Biochem. 94, 223-232), and the pH-dependence curve was interpreted in terms of participation of His 48. On the other hand, the binding constant of the micellar substrate to the modified apoenzyme was much smaller than that for the intact apoenzyme. Nevertheless, the pH-dependence curve could be interpreted in terms of participation of His 48 and Asp 49. From these findings, it was concluded that the ionization state of the alpha-NH2 group of cobra venom phospholipase A2 is essentially irrelevant to the bindings of Ca2+ and also of the monodispersed and micellar substrates.  相似文献   

5.
Phospholipases A2 (PLA2s) are classified into two groups, I and II, according to differences in the polypeptide-chain length and intramolecular disulfide bondings. The hydrolysis of monodispersed 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by bovine pancreas PLA2 (Group I) was studied at 25 degrees C and ionic strength 0.2, and the initial velocity data were analyzed by means of the Michaelis-Menten equation. The Michaelis constant, Km, was found to be practically independent of Ca2+ concentration and also of pH value. The latter result indicates no participation of the ionizable groups in the active site in the substrate binding. The pH-dependence curve of the logarithm of the catalytic center activity, kcat, obtained in the presence of a practically saturating amount of Ca2+, showed three transitions ascribable to the participation of three ionizable groups with pK values of 5.00, 8.40, and 9.50. The respective groups were tentatively assigned to the catalytic group His 48, the N-terminal alpha-amino group, and invariant Tyr 52, which is located in close proximity to the imidazole ring of His 48. Deprotonation of His 48 and protonation of Tyr 52 were shown to be essential to the catalysis. The importance of the ionization state of the alpha-amino group was also indicated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The pH dependence of the binding constant of Ca2+ to a phospholipase A2 of Agkistrodon halys blomhoffii, in which the alpha-amino group had been selectively modified to an alpha-keto group, was studied at 25 degrees C and ionic strength 0.1 by the tryptophyl fluorescence method. The dependence was compared with the results for the intact enzyme (Ikeda et al. (1981) J. Biochem. 90, 1125-1130). The pH-dependence curve could be well interpreted in terms of the participation of the two ionizable groups Asp 49 and His 48, with pK values of 4.70 and 6.69, respectively. These values were slightly different from the respective pK values for the intact enzyme, 5.15 and 6.45. Ca2+ binding to the intact enzyme involves the participation of an additional ionizable group with a pK value of 7.30, which was thus assigned as alpha-amino group. The pH dependence of the binding constant of monodispersed n-dodecylphosphorylcholine (n-C12PC) to the alpha-NH2-modified enzyme was studied at 25 degrees C and ionic strength 0.1 by the aromatic circular dichroism (CD) method. The pH-dependence curve for the modified apoenzyme was interpreted as reflecting the participation of a single ionizable group with a pK value of 4.7, which was assigned to Asp 49 (to which a Ca2+ ion can coordinate) since the curve for the Ca2+ complex lacked this transition: the binding constant was independent of pH. The pH-dependence curves for the intact apoenzyme and its Ca2+ complex involve the participation of an additional ionizable group with pK values of 7.30 and 6.30, respectively (Ikeda & Samejima (1981) J. Biochem. 90, 799-804), which was assigned as the alpha-amino group. The hydrolysis of monodispersed 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by the intact and the alpha-NH2-modified enzymes was studied by the pH stat method at 25 degrees C, pH 8.2, and ionic strength 0.1 in the presence of 3 mM Ca2+. The Km value for the modified enzyme was found to be very similar to that for the intact enzyme: this was compatible with the results of the direct binding study on the monodispersed n-C12PC under the same conditions. However, the kcat value was about 43% of the value for the intact enzyme, suggesting that the alpha-keto group introduced by the chemical modification perturbed the network of hydrogen bonds in the active site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
The hydrolysis of 1,2-dihexanoyl-sn-glycero-3-phosphorylcholine (diC6PC), catalyzed by the phospholipase A2 from the venom of Agkistrodon halys blomhoffii, was studied at 25 degrees C and the ionic strength of 0.1 in the presence of 3-33.3 mM Ca2+, which can saturate the Ca2+-binding site of the enzyme. The initial velocity data, obtained at various concentrations of the substrate below the critical micelle concentration (cmc), were analyzed according to the Michaelis-Menten equation. The pH-dependence curve of the Km value exhibited only one transition below pH 8. The analytical results indicated that the pK value of 6.30 of an ionizable group changed to 6.54 on the binding of the monodispersed substrate. This ionizable group was assigned as the alpha-amino group on the basis of its pK value, which had been determined from the pH dependence of the binding constant of monodispersed n-dodecylphosphorylcholine (n-C12PC) (Ikeda and Samejima (1981) J. Biochem. 90, 799-804, and Haruki et al. (1986) J. Biochem. 99, 99-109). The pH-dependence curve of the kcat value exhibited two transitions, below pH 6.5 and above pH 9.5. The analytical results indicated the participation of two ionizable groups with pK values of 5.55 and 10.50. Deprotonation of the former and protonation of the latter group were found to be essential for the catalysis. The former ionizable group was assigned as His 48 in the active site on the basis of its pK value, which had been determined from the pH dependence of the binding constant of Ca2+ (Ikeda et al. (1981) J. Biochem. 90, 1125-1130).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Phospholipases A2 are classified into two groups, I and II, according to differences in the polypeptide-chain length and the intramolecular-disulfide bondings. The effects of Ca2+ on the kinetic parameters for the hydrolysis of monodispersed and micellar phosphatidylcholines, catalyzed by a cobra (Naja naja atra) enzyme (Group I) and by mamushi (Agkistrodon halys blomhoffii) and habu (Trimeresurus flavoviridis) enzymes (Group II), were studied by the pH-statassay method at 25 degrees C, pH 8.0-8.2, and ionic strength 0.1-0.2. The results were compared with those reported for the other Group I and II enzymes. The Ca2+ binding was clearly shown to be essential for the catalysis of all the phospholipases A2. However, the substrate binding to Group I enzymes was found to be independent of the Ca2+ binding. On the other hand, the substrate binding to Group II enzymes was facilitated more than 10 times by the binding of Ca2+ to the enzymes. This was interpreted in terms of conformation changes of the peptide loop of residues 26 to 44 accompanying the Ca2+ binding. The latter result, but not the former, seems compatible with the hypothesis for interpreting the catalytic mechanism of phospholipases A2 that an intermediate complex should be stabilized by the coordination of the bound Ca2+ ion with the phosphoryl group and the carbonyl oxygen atom of the ester bond at the sn-2 position of the bound substrate molecule [Verheij et al. (1980) Biochemistry 19, 743-750 and (1981) Rev. Physiol. Biochem. Pharmacol. 91, 91-203]. According to the similarity in the primary and tertiary structures of the active sites of both types of enzymes [Renetseder et al. (1985) J. Biol. Chem. 260, 11627-11634], it is supposed that similar intermediate complexes may occur even for Group I enzymes, at least in the transition state of the productive complexes.  相似文献   

9.
Lefkowitz LJ  Deems RA  Dennis EA 《Biochemistry》1999,38(43):14174-14184
Site-directed mutants of the group IA phospholipase A(2) from cobra venom were constructed and expressed in the methylotrophic yeast Pichia pastoris to probe for the proposed phosphatidylcholine (PC) activator site. Previous crystallographic and molecular modeling studies have identified two regions of the enzyme as likely candidates for this site. Residues Glu-55, Trp-61, Tyr-63, Phe-64, and Lys-65 were mutated to test the site advanced by Ortiz et al. [(1992) Biochemistry 31, 2887-2896] while Asp-23 and Arg-30 were mutated to assess the site proposed by Segelke et al. [(1998) J. Mol. Biol. 279, 223-232]. Expressed enzymes were purified by affinity chromatography and analyzed by SDS-PAGE, Western blotting, electrospray ionization mass spectroscopy, and circular dichroism. Both phospholipid headgroup specificity and rates of hydrolysis on monomeric PC substrates were determined and found to be similar for native, wild-type, and all of the mutant enzymes. These results suggest that all of the expressed enzymes were properly folded and contained functional catalytic sites. Mutations of the aromatic residues in the Ortiz site generally had little effect on PC activation, arguing against the importance of this region of the enzyme for PC activation; however, these aromatic amino acids appeared to be important for interfacial activation. In contrast, the D23N mutant in the Segelke site reduced PC activation by 10-fold without affecting activity toward micellar phosphatidylethanolamine substrates. Similar results were found with the D23N/R30M double mutant, suggesting that this region is critical for PC activation. These results provide evidence for the Segelke site as a PC activator site that is distinct from the catalytic site.  相似文献   

10.
Two phospholipase D (PLD) enzymes with both hydrolase and transferase activities were isolated from Streptomyces chromofuscus. There were substantial differences in the kinetic properties of the two PLD enzymes towards monomeric, micellar, and vesicle substrates. The most striking difference was that the higher molecular weight enzyme (PLD57 approximately 57 kDa) could be activated allosterically with a low mole fraction of phosphatidic acid (PA) incorporated into a PC bilayer (Geng et al., J. Biol. Chem. 273 (1998) 12195-12202). PLD42/20, a tightly associated complex of two peptides, one of 42 kDa and the other 20 kDa, had a 4-6-fold higher Vmax toward PC substrates than PLD57 and was not activated by PA. N-Terminal sequencing of both enzymes indicated that both components of PLD42/20 were cleavage products of PLD57. The larger component included the N-terminal segment of PLD57 and contained the active site. The N-terminus of the smaller peptide corresponded to the C-terminal region of PLD57; this peptide had no PLD activity by itself. Increasing the pH of PLD42/20 to 8.9, followed by chromatography of PLD42/20 on a HiTrap Q column at pH 8.5 separated the 42- and 20-kDa proteins. The 42-kDa complex had about the same specific activity with or without the 20-kDa fragment. The lack of PA activation for the 42-kDa protein and for PLD42/20 indicates that an intact C-terminal region of PLD57 is necessary for activation by PA. Furthermore, the mechanism for transmission of the allosteric signal requires an intact PLD57.  相似文献   

11.
Bindings of the phospholipase A2 from Trimeresurus flavoviridis to the monodispersed and micellar n-alkylphosphorylcholines (n-CnPC) were studied at 25 degrees C and ionic strength 0.2 by the aromatic CD and tryptophyl fluorescence methods, respectively. The bindings to micelles of the substrate analog were analyzed by assuming that the micellar surface has multiple binding sites for the enzyme and that these sites are identical and mutually independent. The enzyme binding site was found to accommodate a constant number of the substrate (monomer) molecules, N = 9-13. The binding constant to the micelle was about 40 times greater than it was to the monodispersed substrate. The binding constant to the micellar substrate analog increased on the binding of Ca2+ to the enzyme and decreased on modification of the N-terminal alpha-NH2 group, whereas the binding to the monodispersed substrate analog was independent of pH, of the Ca2+ binding, and of the chemical modification of the alpha-NH2 group. The kinetics of the hydrolyses of monodispersed and micellar dihexanoylphosphatidylcholines (diC6PC) were studied at 25 degrees C and ionic strength 0.2 by the pH-stat method in the presence of saturating amounts of Ca2+. The catalytic center activity, kappa cat, as well as the binding constant, 1/Km, for the micellar substrate, were found to be much greater than those for the monodispersed substrate. The binding constant, 1/Km, of the monodispersed substrate was independent of pH; this was in good agreement with that of the substrate analog described above. The pH-dependence curve of kappa cat for the monodispersed substrate exhibited two transitions, one below pH 6.5 and the other above pH 9.5.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
To elucidate the roles of conserved Asp residues of Bacillus cereus sphingomyelinase (SMase) in the kinetic and binding properties of the enzyme toward various substrates and Mg2+, the kinetic data on mutant SMases (D126G and D156G) were compared with those of wild type (WT) enzyme. The stereoselectivity of the enzyme in the hydrolysis of monodispersed short-chain sphingomyelin (SM) analogs and the binding of Mg2+ to the enzyme were not affected by the replacement of Asp126 or Asp156. The pH-dependence curves of kinetic parameters (1/Km and kcat) for D156G-catalyzed hydrolysis of micellar SM mixed with Triton X-100 (1:10) and of micellar 2-hexadecanoylamino-4-nitrophenylphosphocholine (HNP) were similar in shape to those for WT enzyme-catalyzed hydrolysis. On the other hand, the curves for D126G lacked the transition observed for D156G and WT enzymes. Comparison of the values and the shape of pH-dependence curves of kinetic parameters indicated that Asp126 of WT SMase enhances the enzyme's catalytic activity toward both substrates and its binding of HNP but not SM. The deprotonation of Asp126 enhances the substrate binding and slightly suppresses the catalytic activity toward both substrates. Asp156 of WT SMase acts to decrease the binding of both substrates and the catalytic activity to HNP but not SM. From the present study and the predicted three-dimensional structure of B. cereus SMase, Asp126 was thought to be located close to the active site, and its ionization was shown to affect the catalytic activity and substrate binding.  相似文献   

13.
The phosphodiesterase from snake venom catalyzes the hydrolysis of the Rp diastereomer of thymidine 5'-(4-nitrophenyl [17O,18O]phosphate) in H216O with retention of configuration at phosphorus. This result is in agreement with those previously reported for the hydrolysis of chiral phosphorothioate substrates (Bryant, F. R., and Benkovic, S. J. (1979) Biochemistry 18, 2825-2828; Burgers, P. M. J., Eckstein, F., and Hunneman, D. H. (1979) J. Biol. Chem. 254, 7476-7478). The hydrolysis reaction catalyzed by this enzyme occurs via formation of a covalent nucleotidylated enzyme intermediate.  相似文献   

14.
G Stenberg  P G Board  B Mannervik 《FEBS letters》1991,293(1-2):153-155
Human class Alpha glutathione transferase (GST) A1-1 has been subjected to site-directed mutagenesis of a Tyr residue conserved in all classes of cytosolic GSTs. The change of Tyr8----Phe lowers the specific activities with three substrates to 2-8% of the values for the wild-type enzyme. The changes in the kinetic parameters kcat/KM, Vmax and S0.5 show that the decreased activities are partly due to a reduced affinity for glutathione. The effect is reflected in lowered kcat values, suggesting that the hydroxyl group of Tyr8 is involved in the activation of glutathione. The proposal of such a role for the Tyr residue has support from the 3D structure of a pig lung class Pi GST [Reinemer et al. (1991) EMBO J. 10, 1997-2005]. Thus, Tyr8 appears to be the first active site residue established as participating in the chemical mechanism of a GST.  相似文献   

15.
Yoon MY  Hwang JH  Choi MK  Baek DK  Kim J  Kim YT  Choi JD 《FEBS letters》2003,555(2):185-191
Acetohydroxy acid synthase (AHAS) is one of several enzymes that require thiamine diphosphate and a divalent cation as essential cofactors. Recently, the three-dimensional structure of the enzyme from yeast has been determined [Pang et al., J. Mol. Biol. 317 (2002) 249-262]. While this structure sheds light on the binding of the cofactors and the reaction mechanism, the interactions between the substrates and the enzyme remain unclear. We have studied the pH dependence of kinetic parameters in order to obtain information about the chemical mechanism in the active site. Data are consistent with a mechanism in which substrate selectively catalyzed to the enzyme with an unprotonated base having a pK of 6.48, and a protonated group having a pK of 8.25 for catalysis. The temperature dependence of kinetic parameters was pH-dependent, and the enthalpies of ionization, DeltaH(ion), calculated from the slope of pK(1) and pK(2) are both pH-independent. The solvent perturbation of kinetic parameters was pH-dependent, and the pK(1) from the acidic side and the pK(2) from the basic side were shifted down 0.4 pH units and shifted up 0.6 units as water was replaced by 15% ethanol, respectively. The data are discussed in terms of the acid-base chemical mechanism.  相似文献   

16.
Cajal Y  Berg OG  Jain MK 《Biochemistry》2004,43(28):9256-9264
The interfacial kinetic paradigm is adopted to model the kinetic behavior of pig pancreatic phospholipase A(2) (PLA2) at the monolayer interface. A short delay of about a minute to the onset of the steady state is observed under all monolayer reaction progress conditions, including the PLA2-catalyzed hydrolysis of didecanoylphosphatidyl-choline (PC10) and -glycerol (PG10) monolayers as analyzed in this paper. This delay is independent of enzyme concentration and surface pressure and is attributed to the equilibration time by stationary diffusion of the enzyme added to the stirred subphase to the monolayer through the intervening unstirred aqueous layer. The longer delays of up to several hours, seen with the PC10 monolayers at >15 mN/m, are influenced by surface pressure as well as enzyme concentration. Virtually all features of the monolayer reaction progress are consistent with the assumption that the product accumulates in the substrate monolayer, although the products alone do not spread as a compressible monolayer. These results rule out models that invoke slow "activation" of PLA2 on the monolayer. The observed steady-state rate on monolayers after the delays is <1% of the rate observed with micellar or vesicles substrates of comparable substrate. Together these results suggest that the monolayer steady-state rate includes contributions from steps other than those of the interfacial turnover cycle. Additional considerations that provide understanding of the pre-steady-state behaviors and other nonideal effects at the surface are also discussed.  相似文献   

17.
J D Cortese  S Fleischer 《Biochemistry》1987,26(17):5283-5293
D-beta-Hydroxybutyrate dehydrogenase (BDH) is a lecithin-requiring mitochondrial enzyme that catalyzes the interconversion of beta-hydroxybutyrate and acetoacetate. The purified soluble enzyme devoid of lipid (i.e., the apodehydrogenase) can be reactivated with soluble lecithin or by insertion into phospholipid vesicles containing lecithin. Lipid activation curves have a sigmoidal shape, and two models have been proposed to explain them. We have previously reported that the kinetics of reactivation with short-chain lecithins in the soluble state is consistent with a model in which the enzyme enzyme contains two identical, noninteracting lecithin binding sites, both of which must be occupied to activate the enzyme [noncooperative mechanism; Cortese, J.D., Vidal, J.C., Churchill, P., McIntyre, J.O., & Fleischer, S. (1982) Biochemistry 21, 3899-3908]. More recently a kinetic model involving cooperative interactions between lecithin binding sites was proposed for the reactivation of the membrane-bound enzyme [Sandermann, H., Jr., McIntyre, J.O., & Fleischer, S. (1986) J. Biol. Chem. 261, 6201-6208]. This study reinvestigates the basis for the different conclusions in these two studies. The previous study with soluble lecithins was limited to about 34% of maximal activation compared with mitochondrial phospholipid, due to inactivation of the enzyme at the critical micellar concentration. We could now extend this study to 91% activation by increasing the ethanol concentration. This experimental evidence confirms that the soluble system follows a noncooperative equation. We provide a new kinetic approach to test the cooperative model. A velocity equation is derived for a Hill-type cooperative ligand binding system interacting with a mixture of ligands. This equation predicts a proportionality between an overall weighted cooperative dissociation constant [Kcoop(w)] and a dissociation constant for a single lecithin (PC) species from interacting sites (KPC), regulated by the PC molar fraction (XPC): 1/Kcoop(w) = XPC/KPC. The equation was applied to the data of Sandermann et al. [Sandermann, H., Jr., McIntyre, J.O., & Fleischer, S. (1986) J. Biol. Chem. 261, 6201-6208] as well as to newly obtained data. The results obtained over a wide range of PC molar fractions and different mixtures of bilayer phospholipids fit this equation, confirming the cooperative behavior. We conclude that BDH has a different mode of reactivation depending on the nature of the lipid environment. With soluble lecithin, the activation is noncooperative, whereas in the bilayer, mixtures of phospholipids give cooperative behavior that fits a Hill equation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
A great mystery in the mechanism of phospholipase A2 (PLA2) and many other lipolytic enzymes is the "interfacial activation" induced by micellar but not monomeric substrates. Equally mysterious is the lack of interfacial activation in bee venom PLA2, as opposed to PLA2s from pancreas and other sources. We have probed these problems using the conformationally restricted short-chain cyclopentano-analogues of diacylphosphatidylcholine (Cp-DCnPC, all-trans isomer). In the reaction catalyzed by bovine pancreatic PLA2, Cp-DC8PC behaved differently from DC8PC in that its monomers and micelles showed comparable activities (but lower than the activity of DC8PC). This result suggests that the activity of PLA2 can be regulated by substrate conformation and supports the "substrate conformation model" (Wells, M. A. (1974) Biochemistry 13, 2248-2257), but raises a question as to whether Cp-DC8PC mimics monomers or micelles of DC8PC. Conformational analysis by 1H NMR revealed that monomeric Cp-DC8PC was conformationally restricted near the carbonyl region, a property characteristic of micelles. Thus, monomeric CP-DC8PC can be considered as a conformational analogue of micelles, but the important structural feature lies in the CH2COO region instead of the glycerol backbone. CP-DC8PC was then used to test a previous proposal that the bee venom PLA2 hydrolyzes monomers but not micelles (which would predict little or no activity for Cp-DC8PC since its conformation is micelle-like whether below or above its critical micelle concentration). The results showed that Cp-DC8PC is a relatively good substrate for the bee venom PLA2 in comparison with the pancreatic PLA2. This and other evidence together suggest that the bee venom PLA2 is not sensitive to the conformation of monomeric and micellar substrates and hydrolyzes both monomers and micelles. The results in both PLA2s demonstrate the usefulness of cyclopentano-phospholipids in probing the mechanism of phospholipases and the roles of substrate conformation in the catalysis of PLA2.  相似文献   

19.
F Ghomashchi  B Z Yu  O Berg  M K Jain  M H Gelb 《Biochemistry》1991,30(29):7318-7329
The binding equilibrium of phospholipase A2 (PLA2) to the substrate interface influences many aspects of the overall kinetics of interfacial catalysis by this enzyme. For example, the interpretation of kinetic data on substrate specificity was difficult when there was a significant kinetic contribution from the interfacial binding step to the steady-state catalytic turnover. This problem was commonly encountered with vesicles of zwitterionic phospholipids, where the binding of PLA2 to the interface was relatively poor. The action of PLA2 on phosphatidylcholine (PC) vesicles containing a small amount of anionic phospholipid, such as phosphatidic acid (PA), was studied. It was shown that the hydrolysis of these mixed lipid vesicles occurs in the scooting mode in which the enzyme remains tightly bound to the interface and only the substrate molecules present on the outer monolayer of the target vesicle became hydrolyzed Thus the phenomenon of scooting mode hydrolysis was not restricted to the action of PLA2 on vesicles of pure anionic phospholipids, but it was also observed with vesicles of zwitterionic lipids as long as a critical amount of anionic compound was present. Under such conditions, the initial rate of hydrolysis of PC in the mixed PC/PA vesicles was enhanced more than 50-fold. Binding studies of PLA2 to vesicles and kinetic studies in the scooting mode demonstrated that the enhancement of PC hydrolysis in the PC/PA covesicles was due to the much higher affinity of the enzyme toward covesicles compared to vesicles of pure PC phospholipids. A novel and technically simple protocol for accurate determination of the substrate specificity of PLA2 at the interface was also developed by using a double-radiolabel approach. Here, the action of PLA2 in the scooting mode was studied on vesicles of the anionic phospholipid 1,2-dimyristoyl-sn-glycero-3-phosphomethanol that contained small amounts of 3H- and 14C-labeled phospholipids. From an analysis of the 3H and 14C radioactivity in the released fatty acid products, the ratio of substrate specificity constants (kcat/KMS) was obtained for any pair of radiolabeled substrates. These studies showed that the PLA2s from pig pancreas and Naja naja naja venom did not discriminate between phosphatidylcholine and phosphatidylethanolamine phospholipids or between phospholipids with saturated versus unsaturated acyl chains and that the pig enzyme had a slight preference for anionic phospholipids (2-3-fold). The described protocol provided an accurate measure of the substrate specificity of PLA2 without complications arising from the differences in binding affinities of the enzyme to vesicles composed of pure phospholipids.  相似文献   

20.
Six intramolecularly quenched fluorogenic peptides related to the sequences Phe8 to His13, His6 to His13, and Tyr4 to His13 of the human angiotensinogen, containing o-aminobenzoyl (Abz) and ethylenediamine dinitrophenyl (EDDnp) groups at amino- and carboxyl-terminal amino acids residues, were synthesized by classical solution methods. The Leu-Val is the only bond of all obtained peptides that was hydrolyzed by human renin with different degrees of purity and was resistant to hydrolysis by pig renin and cathepsin D. The hydrolysis of Abz-His-Pro-Phe-His-Leu-Val-Ile-His-EDDnp by human renin was inhibited by a highly specific transition-state analog of angiotensinogen (IC50 = 7.8 x 10(-9) M), described by K. Iizuka et al. (1990, J. Med. Chem. 33, 2707-2714). Therefore, specific and sensitive substrates for the continuous assay of human renin in which as little as 70 microGU of human renin could be detected by Abz-Phe-His-Leu-Val-Ile-His-EDDnp were described. The optimal pHs of hydrolysis of the substrates were in the range 4 to 6.  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