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.
Enzymes that function on membrane surfaces offer many challenges to understanding structural and functional details due to the difficulties of obtaining relevant information of the protein in a physiological environment. Focusing on this aspect of structural biology, it is important to develop conditions that mimic the interaction of membrane proteins with their binding surface and ultimately the mechanisms of action. This approach has been used to characterize the allosteric nature of secreted phospholipase A2 (PLA2) to its substrate interface. The breakthrough here was to crystallize the pancreatic group-IB PLA2 in an anion-assisted dimer with five coplanar phosphate anions bound. In the anion-assisted dimer structure one molecule of a tetrahedral mimic inhibitor and five anions are shared between the two subunits of the dimer. The sn-2-phosphate of the inhibitor, which mimics the tetrahedral intermediate of the esterolysis reaction, is bound in the active site of one subunit, and the alkyl chain extends into the active site slot of the second subunit across the subunit-subunit interface. This interface-bound structural mimic provided insight into the active site environment and specific anionic interactions to the i-face of the protein. The presence or absence of a single critical active site water, corresponds to the difference between the activated or inactivated form of the enzyme. The anion-assisted dimer structure supports a calcium coordinated nucleophilic water mechanism, with its pK(a) modulated by this assisting water. This working model has been further strengthened with an enzyme-product complex structure solved with the hydrolysis products of the substrate PAF also bound to the anion-assisted dimer form of PLA2. Additional confirmation of the assisting-water mechanism comes from a structure of the inactive zymogen proPLA2 also crystallized in an anion-assisted dimer. Remarkably, the assisting water present in the activated complex is absent in this proPLA2 structure.  相似文献   

2.
Evidence is presented that the zymogen of porcine pancreatic phospholipase A2 (prophospholipase A2) interacts with a lipid-water interface provided that the interface has a net negative surface charge. Fluorescence spectroscopy and non-equilibrium gel filtration indicate that binding of prophospholipase A2 (proPLA) to mixed detergent micelles is dependent on the presence of an anionic detergent. Prophospholipase binding is accompanied by a change in the environment of the single tryptophan residue qualitatively similar to that observed when the active enzyme, phospholipase A2 (PLA), binds to micelles. In addition, the rate of tryptic activation of prophospholipase is significantly reduced in the presence of negatively-charged mixed micelles, whereas no change in rate occurs when neutral micelles are present. These observations suggest that the lack of catalytic activity of the zymogen toward organized substrates carrying a negative surface charge cannot be explained by a failure to bind at the lipid-water interface.  相似文献   

3.
We report the structures of the crystallographic dimer of porcine pancreatic IB phospholipase A(2) (PLA2) with either five sulfate or phosphate anions bound. In each structure, one molecule of a tetrahedral mimic MJ33 [1-hexadecyl-3-(trifluoroethyl)-sn-glycero-2-phosphomethanol] and the five anions are shared between the two subunits of the dimer. The sn-2-phosphate of MJ33 is bound in the active site of one subunit (A), and the alkyl chain extends into the active site slot of the second subunit (B) across the subunit-subunit interface. The two subunits are packed together with a large hydrophobic and desolvated surface buried between them along with the five anions that define a plane. The anions bind by direct contact with two cationic residues (R6 and K10) per subunit and through closer-range H-bonding interactions with other polarizable ligands. These features of the "dimer" suggest that the binding of PLA2 to the anionic groups at the anionic interface may be dominated by coordination through H-bonding with only a partial charge compensation needed. Remarkably, the plane defined by the contact surface is similar to the i-face of the enzyme [Ramirez, F., and Jain, M. K. (1991) Proteins: Struct., Funct., Genet. 9, 229-239], which has been proposed to make contact with the substrate interface for the interfacial catalytic turnover. Additionally, these structures not only offer a view of the active PLA2 complexed to an anionic interface but also provide insight into the environment of the tetrahedral intermediate in the rate-limiting chemical step of the turnover cycle. Taken together, our results offer an atomic-resolution structural view of the i-face interactions of the active form of PLA2 associated to an anionic interface.  相似文献   

4.
The changes in the microenvironment of the Trp-3 on the i-face of pig pancreatic IB phospholipase A2 (PLA2) provide a measure of the tight contact (Ramirez and Jain, Protein Sci. 9, 229-239, 1991) with the substrate interface during the processive interfacial turnover. Spectral changes from the single Trp-substituent at position 1, 2, 6, 10, 19, 20, 31, 53, 56 or 87 on the surface of W3F PLA2 are used to probe the Trp-environment. Based on our current understanding only the residue 87 is away from i-face, therefore all other mutants are well suited to report modest differences along the i-face. All Trp-mutants bind tightly to anionic vesicles. Only those with Trp at 1, 2 or 3 near the rim of the active site on the i-face cause significant perturbation of the catalytic functions. Most other Trp-mutants showed < 3-fold change in the interfacial processive turnover rate and the competitive inhibition by MJ33. Binding of calcium to the enzyme in the aqueous phase had modest effect on the Trp-emission intensity. However, on the binding of the enzyme to the interface the fluorescence change is large, and the rate of oxidation of the Trp-substituent with N-bromosuccinimide depends on the location of the Trp-substituent. These results show that the solvation environment of the Trp-substituents on the i-face is shielded in the enzyme bound to the interface. Additional changes are noticeable if the active site of the bound enzyme is also occupied, however, the catalytically inert zymogen of PLA2 (proPLA2) does not show such changes. Significance of these results in relation to the changes in the solvent accessibility and desolvation of the i-face of PLA2 at the interface is discussed.  相似文献   

5.
The changes in the microenvironment of the Trp-3 on the i-face of pig pancreatic IB phospholipase A2 (PLA2) provide a measure of the tight contact (Ramirez and Jain, Protein Sci. 9, 229-239, 1991) with the substrate interface during the processive interfacial turnover. Spectral changes from the single Trp-substituent at position 1, 2, 6, 10, 19, 20, 31, 53, 56 or 87 on the surface of W3F PLA2 are used to probe the Trp-environment. Based on our current understanding only the residue 87 is away from i-face, therefore all other mutants are well suited to report modest differences along the i-face. All Trp-mutants bind tightly to anionic vesicles. Only those with Trp at 1, 2 or 3 near the rim of the active site on the i-face cause significant perturbation of the catalytic functions. Most other Trp-mutants showed < 3-fold change in the interfacial processive turnover rate and the competitive inhibition by MJ33. Binding of calcium to the enzyme in the aqueous phase had modest effect on the Trp-emission intensity. However, on the binding of the enzyme to the interface the fluorescence change is large, and the rate of oxidation of the Trp-substituent with N-bromosuccinimide depends on the location of the Trp-substituent. These results show that the solvation environment of the Trp-substituents on the i-face is shielded in the enzyme bound to the interface. Additional changes are noticeable if the active site of the bound enzyme is also occupied, however, the catalytically inert zymogen of PLA2 (proPLA2) does not show such changes. Significance of these results in relation to the changes in the solvent accessibility and desolvation of the i-face of PLA2 at the interface is discussed.  相似文献   

6.
Pan YH  Yu BZ  Berg OG  Jain MK  Bahnson BJ 《Biochemistry》2002,41(50):14790-14800
We have solved the 1.55 A crystal structure of the anion-assisted dimer of porcine pancreatic group IB phospholipase A2 (PLA2), complexed with the products of hydrolysis of the substrate platelet activating factor. The dimer contains five coplanar phosphate anions bound at the contact surface between the two PLA2 subunits. This structure parallels a previously reported anion-assisted dimer that mimics the tetrahedral intermediate of PLA2 bound to a substrate interface [Pan, Y. H., et al. (2001) Biochemistry 40, 609-617]. The dimer structure has a molecule of the product acetate bound in subunit A and the other product 1-octadecyl-sn-glycero-3-phosphocholine (LPC-ether) to subunit B. Therefore, this structure is of the two individual product binary complexes and not of a ternary complex with both products in one active site of PLA2. Protein crystals with bound products were only obtained by cocrystallization starting from the initial substrate. In contrast, an alternate crystal form was obtained when PLA2 was cocrystallized with LPC-ether and succinate, and this crystal form did not contain bound products. The product bound structure has acetate positioned in the catalytic site of subunit A such that one of its oxygen atoms is located 3.5 A from the catalytic calcium. Likewise, a longer than typical Ca-to-Gly(32) carbonyl distance of 3.4 A results in a final Ca coordination that is four-coordinate and has distorted geometry. The other oxygen of acetate makes hydrogen bonds with N(delta)(1)-His(48), O(delta)(1)-Asp(49), and the catalytic assisting water (w7). In contrast, the glycerophosphocholine headgroup of LPC-ether in subunit B makes no contacts with calcium or with the catalytic residues His(48) or Asp(49). The tail of the LPC-ether is located near the active site pocket with the last nine carbons of the sn-1- acyl chain refined in two alternate conformations. The remaining atoms of the LPC-ether product have been modeled into the solvent channel but have their occupancies set to zero in the refined model due to disorder. Together, the crystallographic and equilibrium binding results with the two products show that the simultaneous binding of both the products in a single active site is not favored.  相似文献   

7.
T Tanaka  S Kimura  Y Ota 《Gene》1988,64(2):257-264
We have developed a phospholipase A2(PLA2)-producing system using Saccharomyces cerevisiae. A 456-bp synthetic DNA fragment was constructed encoding bovine pancreatic phospholipase A2 (proPLA2; zymogen) along with the signal sequence of dog pancreatic PLA2. Yeast-preferred codons were chosen and unique restriction enzyme sites were incorporated. 22 oligodeoxynucleotides that varied in size from 33 to 48 nucleotides were chemically synthesized and assembled into the DNA fragment, which was then placed under the control of the yeast acid phosphatase repressible promoter. The resulting plasmid, transformed into S. cerevisiae, directed the synthesis of about 2.8 micrograms/ml of PLA2, most of which was secreted into the culture fluid. The secreted PLA2 comprised 18 to 26% of active enzyme, the remainder being proenzyme. Both had the expected N-terminal amino acid sequences, indicating that the yeast accurately released the signal peptide and the activation peptide (N-terminal heptapeptide of proPLA2). The specific activity of PLA2 thus produced is the same as that of the authentic bovine enzyme.  相似文献   

8.
We previously reported that PLA(2) activity in the gills is higher than that in other tissues in red sea bream and purified PLA(2) from the gills belongs to the group IB PLA(2) as well as other red sea bream PLA(2)s. In this study, we reconfirmed that the level of PLA(2) activity is extremely high in the gills compared with other tissues, and gill PLA(2) was detected only in the gills by immunoblotting and inhibition test using anti-gill PLA(2) monoclonal antibody. The level of PLA(2) activity and protein expression in the gills are well correlated. Fish can be roughly divided into high and low groups based on the level of PLA(2) activity. Gill PLA(2) was detected in the gills of the high group, but not the low group by immunoblotting. In the gills of the high group, gill PLA(2) was detected in the mucous cells and pavement cells located on the surface of gill epithelia by immunohistochemistry. On the other hand, positive signals were observed only in the mucous cells by in situ hybridization. We also isolated inactive proPLA(2), having AR propeptide, preceding the mature enzyme from the gill extract. These results suggest that gill PLA(2) is synthesized as an inactive proPLA(2) in the mucous cells and is secreted to the surface of gill epithelia.  相似文献   

9.
Fourier transform infrared spectroscopy has been used to investigate the secondary structure of porcine and bovine pancreatic phospholipase A2 (PLA2) and the zymogen of porcine PLA2, prophospholipase A2 (proPLA2), in both H2O and D2O media. Detailed qualitative analysis was made of these proteins using second derivative and deconvolution techniques. Quantitative studies of the proteins in solution made using Factor Analysis gave average values of 54% alpha-helix, 15% beta-sheet and 23% beta-turns. These values agree well with the secondary structures deduced from previous studies of single crystals using X-ray techniques. No significant differences in secondary structure were observed for porcine pancreatic (pro)phospholipase A2 in the presence or absence of Ca2+ ions, or in the temperature range 10-45 degrees C. The binding of the non-degradable phospholipid analogue, n-alkylphosphocholine, in monomeric form produced no significant difference in the secondary structure of either enzyme. Conformational differences were, however, observed between the enzyme lyophilised in a solid film and in aqueous solution. The change is probably due to the formation of beta-sheet upon hydration, coupled with a loss of random structures. Conformational differences in both porcine and bovine pancreatic PLA2 were observed on binding to n-alkylphosphocholine micelles. This change may be due to a small increase in the alpha-helical structure and a decrease in the beta-sheet, and/or possibly beta-turn content. Similar conformational changes were observed for the interaction of porcine and bovine PLA2 with the substrate analogue inhibitor 1-heptanoyl-2-heptanoylamino-2-deoxy-sn-glycero-3-phospho glycol in micellar form.  相似文献   

10.
The cDNA gene encoding porcine pancreatic prophospholipase A2 (proPLA2) was cloned into an Aspergillus niger expression vector downstream of the glucoamylase (glaA) gene promoter region. When this construct was transformed into A. niger, no detectable PLA2 was produced. Evidence was obtained showing that the PLA2 gene was transcribed and that PLA2 is extremely susceptible to both intracellular and extracellular proteases of A. niger, thus indicating that translation products would be rapidly degraded. By fusing the proPLA2-encoding sequence to the entire glaA gene, secreted yields of PLA2 up to 10 micrograms/ml were obtained from a transformed protease-deficient strain of A. niger. PLA2 was secreted in young cultures as a fusion protein, but in older cultures, it was processed from the glucoamylase carrier protein. Secreted PLA2 was shown to be enzymatically active and to have the correct N-terminal amino acid (aa) sequence, although another form of processed PLA2 was also produced. This form included two aa of the proregion from PLA2. The potential for improving yields of secreted heterologous proteins from A. niger still further is discussed.  相似文献   

11.
This is the first crystal structure of a carbohydrate induced dimer of phospholipase A(2) (PLA(2)). This is an endogenous complex formed between two PLA(2) molecules and two mannoses. It was isolated from Krait venom (Bungarus caeruleus) and crystallized as such. The complete amino acid sequence of PLA(2) was determined using cDNA method. Three-dimensional structure of the complex has been solved with molecular replacement method and refined to a final R-factor of 0.192 for all the data in the resolution range 20.0-2.1A. The presence of mannose molecules in the protein crystals was confirmed using dinitrosalicylic acid test and the molecular weight of the dimer was verified with MALDI-TOF. As indicated by dynamic light scattering and analytical ultracentrifugation the dimer was also stable in solution. The good quality non-protein electron density at the interface of two PLA(2) molecules enabled us to model two mannoses. The mannoses are involved extensively in interactions with protein atoms of both PLA(2) molecules. Some of the critical amino acid residues such as Asp 49 and Tyr 31, which are part of the substrate-binding site, are found facing the interface and interacting with mannoses. The structure of the complex clearly shows that the dimerization is caused by mannoses and it results in the loss of enzymatic activity.  相似文献   

12.
Yu BZ  Janssen MJ  Verheij HM  Jain MK 《Biochemistry》2000,39(19):5702-5711
A well-defined region of pancreatic and other secreted phospholipase A2 (PLA2), which we call the i-face, makes a molecular contact with the interface to facilitate and control the events and processivity of the interfacial catalytic turnover cycles. The structural features of the i-face and its allosteric relationship to the active site remain to be identified. As a part of the calcium binding (26-34) loop, Leu-31 is located on the surface near the substrate binding slot of PLA2. Analysis of the primary rate and equilibrium parameters of the Leu-31 substitution mutants of the pig pancreatic PLA2 shows that the only significant effect of the substitution is to impair the chemical step at the zwitterionic interface in the presence of added NaCl, and only a modest effect is seen on kcat at the anionic interface. Leu-31 substitutions have little effect on the binding of the enzyme to the interface; the affinity for certain substrate mimics is modestly influenced in W3F, L31W double mutant. The fluorescence emission results with the double mutant show that the microenvironment of Trp-31 is qualitatively different at the zwitterionic versus anionic interfaces. At both of the interfaces Trp-31 is not shielded from the bulk aqueous environment as it remains readily accessible to acrylamide and water. The NaCl-induced change in the Trp-31 emission spectrum of the double mutant on the zwitterionic interface is similar to that seen on the binding to the anionic interface. Together, the kinetic and spectroscopic results show that the form of PLA2 at the zwitterionic interface (Ez) is distinguishably different from the catalytically more efficient form at the anionic interface (Ea). This finding provides a structural basis for the two-state model for kcat activation by the anionic interface. In conjunction with earlier results we suggest that neutralization of certain cationic residues of PLA2 exerts a control on the calcium loop through residue 31.  相似文献   

13.
Pop C  Feeney B  Tripathy A  Clark AC 《Biochemistry》2003,42(42):12311-12320
The interface of the procaspase-3 dimer plays a critical role in zymogen maturation. We show that replacement of valine 266, the residue at the center of the procaspase-3 dimer interface, with glutamate resulted in an increase in enzyme activity of approximately 60-fold, representing a pseudoactivation of the procaspase. In contrast, substitution of V266 with histidine abolished the activity of the procaspase-3 as well as that of the mature caspase. While the mutations do not affect the dimeric properties of the procaspase, we show that the V266E mutation may affect the formation of a loop bundle that is important for stabilizing the active site. In contrast, the V266H mutation affects the positioning of loop L3, the loop that forms the bulk of the substrate binding pocket. In some cases, the amino acids affected by the mutations are >20 A from the interface. Overall, the results demonstrate that the integrity of the dimer interface is important for maintaining the proper active site conformation.  相似文献   

14.
Caspase‐8 is a cysteine directed aspartate‐specific protease that is activated at the cytosolic face of the cell membrane upon receptor ligation. A key step in the activation of caspase‐8 depends on adaptor‐induced dimerization of procaspase‐8 monomers. Dimerization is followed by limited autoproteolysis within the intersubunit linker (IL), which separates the large and small subunits of the catalytic domain. Although cleavage of the IL stabilizes the dimer, the uncleaved procaspase‐8 dimer is sufficiently active to initiate apoptosis, so dimerization of the zymogen is an important mechanism to control apoptosis. In contrast, the effector caspase‐3 is a stable dimer under physiological conditions but exhibits little enzymatic activity. The catalytic domains of caspases are structurally similar, but it is not known why procaspase‐8 is a monomer while procaspase‐3 is a dimer. To define the role of the dimer interface in assembly and activation of procaspase‐8, we generated mutants that mimic the dimer interface of effector caspases. We show that procaspase‐8 with a mutated dimer interface more readily forms dimers. Time course studies of refolding also show that the mutations accelerate dimerization. Transfection of HEK293A cells with the procaspase‐8 variants, however, did not result in a significant increase in apoptosis, indicating that other factors are required in vivo. Overall, we show that redesigning the interface of procaspase‐8 to remove negative design elements results in increased dimerization and activity in vitro, but increased dimerization, by itself, is not sufficient for robust activation of apoptosis.  相似文献   

15.
In a recent series of kinetic studies (De Haas et al. (1990) Biochim. Biophys. Acta 1046, 249-257 and references therein) we have demonstrated that synthetic (R)-phospholipid analogues containing a 2-acylaminogroup instead of the 2-acyloxy function found in natural phospholipids, behave as strong competitive inhibitors of porcine pancreatic phospholipase A2 (PLA2). We also showed that these analogues strongly bind to the active site of the enzyme but only after their incorporation into a micellar substrate/water interface. In the present study we investigated the interaction of native PLA2 and of an inactive PLA2 in which the active site residue His-48 has been modified by alkylation with 1-bromo-2-octanone, with pure micelles of several of these inhibitors in both enantiomeric forms by means of ultraviolet difference absorption spectroscopy. Our results show that the first interaction step between native or modified enzyme and micellar lipid/water interfaces probably consists of a low-affinity Langmuir-type adsorption characterized by signals arising from the perturbation of the single Trp-3 residue. Once present at the interface the native enzyme is able to bind, in a second step, a single inhibitor molecule of the (R)-configuration in its active site, whereas the (S)-enantiomer is not bound in the active site. The overall dissociation constant of the interfacial phospholipase-inhibitor complex is three orders of magnitude lower for micelles composed of the (R)-isomer than those of the (S)-isomer. The modified PLA2 still adsorbs to micellar lipid/water interfaces but cannot bind either of the two enantiomers into its active site and similar dissociation constants were found for lipid-protein complexes with micelles of either the (R) or the (S) inhibitors. After blanking the ultraviolet signals due to the perturbation of Trp-3 in the initial adsorption step of the enzyme to a micellar surface of a non-inhibitory phospholipid analogue, the progressive binding of a single (R)-inhibitor molecule into the active site could be followed quantitatively by a tyrosine perturbation. These titrations yielded numerical values for the dissociation constants in the interface and provide a possible explanation for the large difference in overall dissociation constants of the complexes between enzyme and micelles of (R)-and (S)-inhibitors. With the use of PLA2 mutants in which each time a single tyrosine was replaced by phenylalanine, the tyrosine residues involved in binding of the monomeric inhibitor molecule were identified as Tyr-69 and Tyr-52.  相似文献   

16.
In the present study, three Taiwan cobra PLA(2) variants were prepared by adding an extra N-terminal Met, substituting Asn-1 by Met or deleting the N-terminal heptapeptide. Recombinant PLA(2) mutants were expressed in Escherichia coli (E. coli), and purified to homogeneity by reverse phase HPLC. Fluorescence measurement showed that the hydrophobic character of the catalytic site, the microenvironment of Trp residues and energy transfer from excited Trp to 8-anilinonaphthalene sulfonate (ANS) were affected by N-terminal mutations. An alteration in the structural flexibility of the active site was noted with the mutants lacking the N-terminal heptapeptide or with an extra N-terminal Met added as evidenced by the inability of the two variants to bind with Ba(2+). Moreover, modification of Lys residues and energy transfer within the protein-ANS complex revealed that the Ca(2+)-induced change in the global structure of PLA(2) was different from that in N-terminal variants. Together with the fact that an 'activation network' connects the N-terminus with the active site, our data suggest that mutagenesis on the N-terminal region affects directly the fine structure of the catalytic site, which subsequently transmits its influence in altering the structure outside the active site of PLA(2). Copyright (c) 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
The cDNA coding for the porcine pancreatic prophospholipase A2 (proPLA) has been cloned and expressed in E. coli. Expression of proPLA could only be obtained in the form of intracellular aggregates after fusing the 15 kDa proPLA to a large (greater than or equal to 45 kDa) bacterial peptide. The fusion protein was readily purified from cell lysates, and specifically cleaved. Cleavage of the fusion protein was achieved with either hydroxylamine (at Asn/Gly sequences in the denatured protein), or trypsin (between the pro- and the mature PLA in the renatured fusion protein). The former method releases a proPLA-like enzyme, while the latter directly yields PLA. Renaturation of the fusion protein was made possible by the use of a recently reported new S-sulphonation method. The released (pro)PLA was purified (yields of 2-3 mg/ltr of culture medium), and showed identical properties compared to native (pro)PLA.  相似文献   

18.
The negatively charged detergents S-n-alka-noylthioglycol sulfates (C8, C9, and C10) are substrates for porcine pancreatic phospholipase A2 and its zymogen. At pH 6.0 and detergent concentrations up to 0.08 X critical micelle concentration (cmc), the activities of active enzyme and zymogen are similar and very low. From 0.08 X cmc to 0.12 X cmc a tremendous increase in activity is observed for phospholipase A2, but not for the zymogen. Concomitant with this increase in activity there is a sharp rise in molecular weight of the substrate-enzyme complex, from 15 000 to 95 000, and in detergent to protein molar ratio of 1:1 to about 7:1. This indicates both substrate and enzyme aggregation. Most probably a lipid-water interface is formed inside the aggregated protein particle by which the enzyme is activated. Although the zymogen also forms high molecular weight complexes with similar molar ratios, no activation is observed probably because of distortion of its lipid binding domain.  相似文献   

19.
A myotoxic Asp49-phospholipase A2 (Asp49-PLA2) with low catalytic activity (BthTX-II from Bothrops jararacussu venom) was crystallized and the molecular-replacement solution has been obtained with a dimer in the asymmetric unit. The quaternary structure of BthTX-II resembles the myotoxic Asp49-PLA2 PrTX-III (piratoxin III from B. pirajai venom) and all non-catalytic and myotoxic dimeric Lys49-PLA2S. Despite of this, BthTX-II is different from the highly catalytic and non-myotoxic BthA-I (acidic PLA2 from B. jararacussu) and other Asp49-PLA2S. BthTX-II structure showed a severe distortion of calcium-binding loop leading to displacement of the C-terminal region. Tyr28 side chain, present in this region, is in an opposite position in relation to the same residue in the catalytic activity Asp49-PLA2S, making a hydrogen bond with the atom O delta 2 of the catalytically active Asp49, which should coordinate the calcium. This high distortion may also be confirmed by the inability of BthTX-II to bind Na+ ions at the Ca2+-binding loop, despite of the crystallization to have occurred in the presence of this ion. In contrast, other Asp49-PLA2S which are able to bind Ca2+ ions are also able to bind Na+ ions at this loop. The comparison with other catalytic, non-catalytic and inhibited PLA2S indicates that the BthTX-II is not able to bind calcium ions; consequently, we suggest that its low catalytic function is based on an alternative way compared with other PLA2S.  相似文献   

20.
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   

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