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.
Berg OG  Yu BZ  Chang C  Koehler KA  Jain MK 《Biochemistry》2004,43(25):7999-8013
Equilibrium parameters for the binding of monodisperse alkyl sulfate along the i-face (the interface binding surface) of pig pancreatic IB phospholipase A(2) (PLA2) to form the premicellar complexes (E(i)(#)) are characterized to discern the short-range specific interactions. Typically, E(i)(#) complexes are reversible on dilution. The triphasic binding isotherm, monitored as the fluorescence emission from the single tryptophan of PLA2, is interpreted as a cooperative equilibrium for the sequential formation of three premicellar complexes (E(i)(#), i = 1, 2, 3). In the presence of calcium, the dissociation constant K(1) for the E(1)(#) complex of PLA2 with decyl sulfate (CMC = 4500 microM) is 70 microM with a Hill coefficient n(1) = 2.1 +/- 0.2; K(2) for E(2)(#) is 750 microM with n(2) = 8 +/- 1, and K(3) for E(3)(#) is 4000 microM with an n(3) value of about 12. Controls show that (a) self-aggregation of decyl sulfate alone is not significant below the CMC; (b) occupancy of the active site is not necessary for the formation of E(i)(#); (c) K(i) and n(i) do not change significantly due to the absence of calcium, possibly because alkyl sulfate does not bind to the active site of PLA2; (d) the E(i)(#) complexes show a significant propensity for aggregation; and (e) PLA2 is not denatured in E(i)(#). The results are interpreted to elaborate the model for atomic level interactions along the i-face: The chain length dependence of the fit parameters suggests that short-range specific anion binding of the headgroup is accompanied by desolvation of the i-face of E(i)(#). We suggest that allosteric activation of PLA2 results from such specific interactions of the amphiplies and the desolvation of the i-face. The significance of these primary interfacial binding events and the coexistence of the E and E(i)(#) aggregates is discussed.  相似文献   

2.
Bai S  Jain MK  Berg OG 《Biochemistry》2008,47(9):2899-2907
Pig pancreatic IB phospholipase A 2 (PLA2) forms three distinguishable premicellar E i (#) ( i = 1, 2, and 3) complexes at successively higher decylsulfate concentrations. The Hill coefficient for E 1 (#) is n 1 = 1.6, and n 2 and n 3 for E 2 (#) and E 3 (#) are about 8 each. Saturation-transfer difference nuclear magnetic resonance (NMR) and other complementary results with PLA2 show that decylsulfate molecules in E 2 (#) and E 3 (#) are contiguously and cooperatively clustered on the interface-binding surface or i-face that makes contact with the substrate interface. In these complexes, the saturation-transfer difference NMR signatures of (1)H in decylsulfate are different. The decylsulfate epitope for the successive E i (#) complexes increasingly resembles the micellar complex formed by the binding of PLA2 to preformed micelles. Contiguous cooperative amphiphile binding is predominantly driven by the hydrophobic effect with a modest electrostatic shielding of the sulfate head group in contact with PLA2. The formation of the complexes is also associated with structural change in the enzyme. Calcium affinity of E 2 (#) appears to be modestly lower than that of the free enzyme and E 1 (#). Binding of decylsulfate to the i-face does not require the catalytic calcium required for the substrate binding to the active site and for the chemical step. These results show that E i (#) complexes are useful to structurally characterize the cooperative sequential and contiguous binding of amphiphiles on the i-face. We suggest that the allosteric changes associated with the formation of discrete E i (#) complexes are surrogates for the catalytic and allosteric states of the interface activated PLA2.  相似文献   

3.
Berg OG  Yu BZ  Apitz-Castro RJ  Jain MK 《Biochemistry》2004,43(7):2080-2090
Phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus cereus forms a premicellar complex E(#) with monodisperse diheptanoylphosphatidylcholine (DC(7)PC) that is distinguishable from the E complex formed with micelles. Results are interpreted with the assumption that in both cases amphiphiles bind to the interfacial binding surface (i-face) of PI-PLC but not to the active site. Isothermal calorimetry and fluorescence titration results for the binding of monodisperse DC(7)PC give an apparent dissociation constant of K(2) = 0.2 mM with Hill coefficient of 2. The gel-permeation, spectroscopic, and probe partitioning behaviors of E(#) are distinct from those of the E complex. The aggregation and partitioning behaviors suggest that the acyl chains in E(#) but not in E remain exposed to the aqueous phase. The free (E) and complexed (E(#) and E) forms of PI-PLC, each with distinct spectroscopic signatures, readily equilibrate with changing DC(7)PC concentration. The underlying equilibria are modeled and their significance for the states of the PI-PLC under monomer kinetic conditions is discussed to suggest that the Michaelis-Menten complex formed with monodisperse DC(7)PC is likely to be E(#)S or its aggregate rather than the classical monodisperse ES complex.  相似文献   

4.
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   

5.
The structural properties of bacteriophage M13 during disassembly were studied in different membrane model systems, composed of a homologue series of the detergents sodium octyl sulfate, sodium decyl sulfate, and sodium dodecyl sulfate. The structural changes during phage disruption were monitored by spin-labeled electron spin resonance (ESR) and circular dichroism spectroscopy. For the purpose of ESR spectroscopy the major coat protein mutants V31C and G38C were site-directed spin labeled in the intact phage particle. These mutants were selected because the mutated sites are located in the hydrophobic part of the protein, and provide good reporting locations for phage integrity. All amphiphiles studied were capable of phage disruption. However, no significant phage disruption was detected below the critical micelle concentration of the amphiphile used. Based on this finding and the linear dependence of phage disruption by amphiphiles on the phage concentration, it is suggested that the solubilization of the proteins of the phage coat by amphiphiles starts with an attachment to and penetration of amphiphile molecules into the phage particle. The amphiphile concentration in the phage increases in proportion to the amphiphile concentration in the aqueous phase. Incorporation of the amphiphile in the phage particle is accompanied with a change in local mobility of the spin-labeled part of the coat protein and its secondary structure. With increasing the amphiphile concentration in the phage particle, a concentration is reached where the concentration of the amphiphile in the aqueous phase is around its critical micelle concentration. A further increase in amphiphile concentration results in massive phage disruption. Phage disruption by amphiphiles appears to be dependent on the phage coat mutations. It is concluded that phage disruption is dependent on a hydrophobic effect, since phage solubilization could significantly be increased by keeping the hydrophilic part of the amphiphile constant, while increasing its hydrophobic part.  相似文献   

6.
Phospholipase A2 (PLA2) activities were found in the cytosolic fractions of rat brain. Using the gel filtration chromatography, two major peaks of PLA2 activities were demonstrated: PLA2-H (200-500 kDa) and PLA2-L (100 kDa). PLA2-L was active at both neutral and alkaline pH and absolutely required Ca2+ for the activity, while the activity of PLA2-H was detected only at alkaline pH and independent of Ca2+. The activation of PLA2-L by Ca2+ was biphasic; the first observed at 1-100 microM Ca2+ and the second at 10 mM Ca2+. In the reconstitution system of partially purified PLA2-L and synaptosomal membranes from rat brain, PLA2-L associated with the membranes in a Ca2(+)-dependent manner. The association was completed within 5-10 min at 25 degrees C both at 10 microM and 1 mM Ca2+, though amount of PLA2-L translocated was dependent on Ca2+ concentrations. These results suggest that Ca2+ promotes the translocation of the cytosolic PLA2-L to membranes where phospholipids, substrate of PLA2, are present.  相似文献   

7.
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   

8.
Two carboxylesterases (ME-III and ME-IV) have been purified to apparent homogeneity from the seeds of Mucuna pruriens employing ammonium sulfate fractionation, cation exchange chromatography on CM-cellulose, gel-permeation chromatography on Sephadex G-100 and preparative PAGE. The homogeneity of the purified preparations was confirmed by polyacrylamide gel electrophoresis (PAGE), gel-electrofocussing and SDS–PAGE. The molecular weights determined by gel-permeation chromatography on Sephadex G-200 were 20.89 kDa (ME-III) and 31.62 kDa (ME-IV). The molecular weights determined by SDS–PAGE both in the presence and absence of 2-mercaptoethanol were 21 kDa (ME-III) and 30.2 kDa (ME-IV) respectively, suggesting a monomeric structure for both the enzymes. The enzymes were found to have Stokes radius of 2.4 nm (ME-III) and 2.7 nm (ME-IV). The isoelectric pH values of the enzymes, ME-III and ME-IV, were 6.8 and 7.4, respectively. ME-III and ME-IV were classified as carboxylesterases employing PAGE in conjunction with substrate and inhibitor specificity. The Km of ME-III and ME-IV with 1-naphthyl acetate as substrate was 0.1 and 0.166 mM while with 1-naphthyl propionate as substrate the Km was 0.052 and 0.0454 mM, respectively. As the carbon chain length of the acyl group increased, the affinity of the substrate to the enzyme increased indicating hydrophobic nature of the acyl group binding site. The enzymes exhibited an optimum temperature of 45 °C (ME-III) and 37 °C (ME-IV), an optimum pH of 7.0 (ME-III) and 7.5 (ME-IV) and both the enzymes (ME-III and ME-IV) were stable up to 120 min at 35 °C. Both the enzymes were inhibited by organophosphates (dichlorvos and phosphamidon), but resistant towards carbamates (carbaryl and eserine sulfate) and sulphydryl inhibitors (p-chloromercuricbenzoate, PCMB).  相似文献   

9.
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   

10.
For membrane-associated enzymes, which access substrate from either a monolayer or bilayer of the aggregate substrate, the partitioning from the aqueous phase to this phospholipid interface is critical for catalysis. Despite a large and expanding body of knowledge regarding interfacial enzymes, the biophysical steps involved in interfacial recognition and adsorption remain relatively poorly understood. The surface of the enzyme that contacts the phospholipid surface is referred to as its interfacial binding surface, or more simply, its i-face. The interaction of a protein's i-face with the aggregate substrate may simply control access to substrate. However, it can be more complex, and this interaction often serves to allosterically activate the enzyme on this surface. First we briefly review what is currently known about i-face structure and function for a prototypical interfacial enzyme, the secreted Phospholipase A2 (PLA2). Then we develop, characterize, compare, and discuss models of the PLA2 i-face across a subset of five homologous PLA2 family members, groups IA, IB, IIA, V, and X. A homology model of human group-V is included in this comparison, suggesting that a similar approach could be used to explore interfacial function of any of the PLA2 family members. Despite moderate sequence identity, structural homology and sequence similarity are well conserved. We find that the residues predicted to be interfacial, while conserved structurally, are not highly conserved in sequence. Implications for this divergence on interfacial selectivity are discussed.  相似文献   

11.
Using defatted and SH-blocked bovine serum albumin (BSA), measurements of differential scanning calorimetry (DSC) have been made at pH 7 on the complexes of BSA and a series of sodium alkyl sulfates. Alkyl sulfates used were sodium decyl sulfate (SDeS), sodium octyl sulfate (SOS), sodium hexyl sulfate (SHS) and sodium ethyl sulfate (SES). Results obtained were compared with those on the system BSA-sodium dodecyl sulfate (SDS) studied previously. Two peaks P 1 and P2 existed in the DSC curve of BSA. These peaks originate in the heat-induced transition of BSA. The pattern of DSC curve changed with the amount of the ligand added, i.e. with the molar mixing ratio ligand/BSA (1). The change for systems BSA-SDeS, BSA-SOS and BSA-SHS was qualitatively the same as that for the system BSA-SDS (2). Interestingly, SES, which is not a surfactant, interacts with BSA. The change for the system BSA-SES was qualitatively the same as that for the system BSA-Na2SO4. All alkyl sulfates suppressed the heat-induced transition at lower concentrations. A linear relationship was obtained for the plots of log(D/A)1 versus log CMC, where (D/A)1 is the molar mixing ratio of anionic surfactant (D) to BSA (A) at which the most heat-stable complex is formed. This suggests that the hydrophobic force has a serious effect on the formation of heat-stable complexes.  相似文献   

12.
An enzyme with sulfatase activity has been isolated from the granules of a rat NK leukemia cell line, CRNK-16. The enzyme has been purified from crude preparation, with a specific activity of 52 nmol/min/mg of protein, by DEAE ion exchange and Con A-Sepharose affinity chromatography, resulting in a specific activity of 230 nmol/min/mg of protein. The molecular mass of the purified enzyme was estimated to be 40 kDa by gel filtration chromatography at pH 7.4, but the enzyme had the ability to complex to molecular masses of greater than 300 kDa at low pH when crude granule extract was used as the starting sample, suggesting that it associates with other granule components. The enzyme was determined to be an arylsulfatase by its ability to (a) hydrolyze p-nitrophenyl sulfate (Km = 26.0 mM) and p-nitrocatechol sulfate (pNC sulfate) (Km = 1.1 mM) and (b) be inhibited by sulfite (Ki = 6.0 x 10(-7) M), sulfate (Ki = 1 x 10(-3) M), and phosphate (Ki = 4 x 10(-5) M) in a competitive manner. The pH optimum for enzymatic activity was determined to be 5.6. The role of this enzyme in cytolytic function was investigated by examining the effect of its substrates and inhibitors on granule- and cell-mediated lysis. pNC sulfate was shown to cause a dose-dependent inhibition of target cell lysis by isolated cytolytic granules (complete inhibition at 12.5 mM). Sulfite induced an incomplete inhibition (50% at 1 mM), whereas phosphate was essentially without inhibitory effect. Sulfate, on the other hand, altered lytic activity in a biphasic manner, inasmuch as it induced an inhibition of lysis at high concentrations and an increase of lysis at low concentrations. Cell-mediated lysis was inhibited by pNC sulfate in a dose-dependent fashion at concentrations greater than 2.5 mM, with nearly complete inhibition at 50 mM. Sulfate also altered the lytic activity by intact cells in a biphasic manner, although the effect was much less pronounced. Sulfite and phosphate caused only a 30% inhibition of lytic activity. These results suggest that the sulfatase enzyme is involved in NK cytolytic function, presumably at the lethal hit stage.  相似文献   

13.
During the steady state reaction progress in the scooting mode with highly processive turnover, Bacillus cereus sphingomyelinase (SMase) remains tightly bound to sphingomyelin (SM) vesicles (Yu et al., Biochim. Biophys. Acta 1583, 121-131, 2002). In this paper, we analyze the kinetics of SMase-catalyzed hydrolysis of SM dispersed in diheptanoylphosphatidyl-choline (DC7PC) micelles. Results show that the resulting decrease in the turnover processivity induces the stationary phase in the reaction progress. The exchange of the bound enzyme (E*) between the vesicle during such reaction progress is mediated via the premicellar complexes (E(i)#) of SMase with DC7PC. Biophysical studies indicate that in E(i)# monodisperse DC7PC is bound to the interface binding surface (i-face) of SMase that is also involved in its binding to micelles or vesicles. In the presence of magnesium, required for the catalytic turnover, three different complexes of SMase with monodisperse DC7PC (E(i)# with i=1, 2, 3) are sequentially formed with Hill coefficients of 3, 4 and 8, respectively. As a result, during the stationary phase reaction progress, the initial rate is linear for an extended period and all the substrate in the reaction mixture is hydrolyzed at the end of the reaction progress. At low mole fraction (X) of total added SM, exchange is rapid and the processive turnover is limited by the steps of the interfacial turnover cycle without becoming microscopically limited by local substrate depletion or enzyme exchange. At high X, less DC7PC will be monodisperse, E(i)# does not form and the turnover becomes limited by slow enzyme exchange. Transferred NOESY enhancement results show that monomeric DC7PC in solution is in a rapid exchange with that bound to E(i)# at a rate comparable to that in micelles. Significance of the exchange and equilibrium properties of the E(i)# complexes for the interpretation of the stationary phase reaction progress is discussed.  相似文献   

14.
Structural basis for bile salt inhibition of pancreatic phospholipase A2   总被引:1,自引:0,他引:1  
Bile salt interactions with phospholipid monolayers of fat emulsions are known to regulate the actions of gastrointestinal lipolytic enzymes in order to control the uptake of dietary fat. Specifically, on the lipid/aqueous interface of fat emulsions, the anionic portions of amphipathic bile salts have been thought to interact with and activate the enzyme group-IB phospholipase A2 (PLA2) derived from the pancreas. To explore this regulatory process, we have determined the crystal structures of the complexes of pancreatic PLA2 with the naturally occurring bile salts: cholate, glycocholate, taurocholate, glycochenodeoxycholate, and taurochenodeoxycholate. The five PLA2-bile salt complexes each result in a partly occluded active site, and the resulting ligand binding displays specific hydrogen bonding interactions and extensive hydrophobic packing. The amphipathic bile salts are bound to PLA2 with their polar hydroxyl and sulfate/carboxy groups oriented away from the enzyme's hydrophobic core. The impaired catalytic and interface binding functions implied by these structures provide a basis for the previous numerous observations of a biphasic dependence of the rate of PLA2 catalyzed hydrolysis of zwitterionic glycerophospholipids in the presence of bile salts. The rising or activation phase is consistent with enhanced binding and activation of the bound PLA2 by the bile salt induced anionic charge in a zwitterionic interface. The falling or inhibitory phase can be explained by the formation of a catalytically inert stoichiometric complex between PLA2 and any bile salts in which it forms a stable complex. The model provides new insight into the regulatory role that specific PLA2-bile salt interactions are likely to play in fat metabolism.  相似文献   

15.
The polyamine, spermine (1-5 mM), when added to rat thyroid cytosol, increases the phosphorylation of a 107 kDa protein 4-fold as analyzed by sodium dodecyl sulfate polyacrylamide gradient gel electrophoresis (SDS-PAGE) and autoradiography; spermidine was less effective and putrescine was without effect. Sodium chloride, when tested at equivalent ionic strengths (4-40 mM), did not reproduce the effects of spermine. In addition to stimulating the phosphorylation of a 107 kDa protein, spermine had an apparent biphasic effect on the phosphorylation of 88 and 65 kDa proteins; maximum stimulation of approximately 60-70% was observed at 0.5-2 mM. Both basal and spermine-stimulated protein phosphorylation patterns were identical whether [gamma-32P] ATP or [gamma-32P] GTP was used as phosphate donors. Heparin (1 microgram/ml) reduced spermine-stimulated phosphorylation of the 107 kDa protein by 64%. Phosphorylation of a 107 kDa protein was not restricted to rat thyroid as spermine was found to augment the phosphorylation of 107 kDa protein(s) in mouse and beef thyroid cytosol preparations.  相似文献   

16.
In order to evaluate the differences in the partition properties of 35 structurally congeneric nucleobases of biological interests in octanol-water biphasic, alkyl C(8)/C(18), and IAM systems, a comparative chromatographic study was performed. Comparing with the reversed-phase C(8)/C(18) retention data, most of the purines possessed weaker IAM retention except for those with specific H-bond and/or electrostatic interactions. Quantitative correlations between the experimental log P(ow) literature values and the IAM, C(8), and C(18) log k were evaluated (R(2)=0.943, 0.794, and 0.767, respectively). Although IAM retention correlated significantly better (larger R(2) value) with the log P(ow) values statistically, the latter was revealed apparently behaving more like (slope approaching unity) alkyl C(8)/C(18) retention and hence also has the same shortcoming in under-representing analytes capable of forming short-term H-bond/electrostatic interactions with polar head-groups of phospholipids. A chemically meaningful structure-retention model (q(2)=0.824 and R(2)=0.968) was derived, in which the hydrophobic interaction is identified as the underlying factor for the retention of purines in IAM system modulated non-trivially by H-bond/electrostatic interactions.  相似文献   

17.
The properties of three discrete premicellar complexes (E1#, E2#, E3#) of pig pancreatic group-IB secreted phospholipase A2 (sPLA2) with monodisperse alkyl sulfates have been characterized [Berg, O. G. et al., Biochemistry 43, 7999–8013, 2004]. Here we have solved the 2.7 Å crystal structure of group-IB sPLA2 complexed with 12 molecules of octyl sulfate (C8S) in a form consistent with a tetrameric oligomeric that exists during the E1# phase of premicellar complexes. The alkyl tails of the C8S molecules are centered in the middle of the tetrameric cluster of sPLA2 subunits. Three of the four sPLA2 subunits also contain a C8S molecule in the active site pocket. The sulfate oxygen of a C8S ligand is complexed to the active site calcium in three of the four protein active sites. The interactions of the alkyl sulfate head group with Arg-6 and Lys-10, as well as the backbone amide of Met-20, are analogous to those observed in the previously solved sPLA2 crystal structures with bound phosphate and sulfate anions. The cluster of three anions found in the present structure is postulated to be the site for nucleating the binding of anionic amphiphiles to the interfacial surface of the protein, and therefore this binding interaction has implications for interfacial activation of the enzyme.  相似文献   

18.
S Chakraborti  J R Michael 《FEBS letters》1991,281(1-2):185-187
Exposure of bovine pulmonary arterial endothelial cells to 1 mM H2O2 stimulated associated TAME-esterase and PLA2 activities. Pretreatment with the serine esterase inhibitors: PMSF (1 mM), DFP (1 mM), and alpha 1-PI (1 mg/ml) inhibited H2O2-induced stimulation of TAME-esterase and PLA2 activities. The TAME-esterase and PLA2 activities under H2O2 exposure were determined to be linearly correlated. Affinity labelling of the endothelial cell membrane with [3H]DFP demonstrated that the serine esterase resides in a protein having molecular weight of 29,000 daltons (29 kDa) which is similar to that of elastase. Treatment of the endothelial cell homogenate with trypsin (1 microgram/ml) also stimulated PLA2 activity.  相似文献   

19.
Cultured arterial smooth muscle cells synthesize a cell-associated heparan sulfate proteoglycan which consists of a 92 kDa core protein with 3 to 4 heparan sulfate side chains covalently attached. Biosynthesis of the cell-associated heparan sulfate proteoglycan was compared in proliferating and in non-dividing vascular smooth muscle cells which are preincubated in the presence of [35]sulfate or a combination of [35S]methionine and [3H]glucosamine. The Mr of the core protein was identical in either growth state, but changes in the structure of the heparan sulfate side chains were observed. Non-dividing (postconfluent) arterial smooth muscle cells form longer heparan sulfate chains with a higher proportion of hydrophobic (N-acetyl) groups than proliferating (preconfluent) cells as judged from gel filtration experiments, hydrophobic interaction chromatography and heparitinase degradation. An enzyme preparation from proliferating cells catalyzes deacetylation and N-sulfation of heparan sulfate at a 5-fold higher activity than from non-dividing cells. Cell density-dependent structural differences of heparan sulfate are related to the finding that heparan sulfate isolated from non-dividing cells has a 10-fold higher antiproliferative potency than heparan sulfate from proliferating (preconfluent) cells.  相似文献   

20.
To evaluate the role of hydrophobic and electrostatic or other polar interactions for protein–ligand binding, we have studied the interactions of bovine serum albumin (BSA) with 2-alkylmalonic acid and 2-alkylbenzimidazole amphiphiles having different head group and alkyl chain length. The binding affinity for the protein–amphiphile interactions is found to depend predominantly on the length of hydrocarbon chain, suggesting the crucial role of hydrophobic forces, supported by polar interactions at the protein surface. The BSA fluorescence exhibits appreciable hypsochromic shift along with a reduction in fluorescence intensity and mean lifetime upon binding with 2-alkylmalonic acid. UV–visible, steady state and time-resolved fluorescence measurements were performed to compare the effects of amphiphiles on BSA as a function of the amphiphiles head group and alkyl chain length.  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