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.
Enzyme inhibition studies on phosphatidylinositol-specific phospholipase C (PI-PLC) from B. Cereus were performed in order to gain an understanding of the mechanism of the PI-PLC family of enzymes and to aid inhibitor design. Inhibition studies on two synthetic cyclic phosphonate analogues (1,2) of inositol cyclic-1:2-monophosphate (cIP), glycerol-2-phosphate and vanadate were performed using natural phosphatidylinositol (PI) substrate in Triton X100 co-micelles and an NMR assay. Further inhibition studies on PI-PLC from B. Cereus were performed using a chromogenic, synthetic PI analogue (DPG-PI), an HPLC assay and Aerosol-OT (AOT), phytic acid and vanadate as inhibitors. For purposes of comparison, a model PI-PLC enzyme system was developed employing a synthetic Cu(II)-metallomicelle and a further synthetic PI analogue (IPP-PI). The studies employing natural PI substrate in Triton X100 co-micelles and synthetic DPG-PI in the absence of surfactant indicate three classes of PI-PLC inhibitors: (1) active-site directed inhibitors (e.g. 1,2); (2) water-soluble polyanions (e.g. tetravanadate, phytic acid); (3) surfactant anions (e.g. AOT). Three modes of molecular recognition are indicated to be important: (1) active site molecular recognition; (2) recognition at an anion-recognition site which may be the active site, and; (3) interfacial (or hydrophobic) recognition which may be exploited to increase affinity for the anion-recognition site in anionic surfactants such as AOT. The most potent inhibition of PI-PLC was observed by tetravanadate and AOT. The metallomicelle model system was observed to mimic PI-PLC in reproducing transesterification of the PI analogue substrate to yield cIP as product and in showing inhibition by phytic acid and AOT.  相似文献   

2.
The deacylated forms of the phosphoinositides were used to determine whether the guinea pig uterus phosphoinositide-specific phospholipase C (PI-PLC I, Mr 60,000) required fatty acids at the sn-1 and sn-2 positions for the hydrolysis of the sn-3 phosphodiester bond. L-alpha-Glycerophospho-D-myo-inositol 4-phosphate (Gro-PIP), but not glycerol 3-phosphate (Gro-3-P), L-alpha-glycerophospho-D-myo-inositol (Gro-PI), or L-alpha-glycerophospho-D-myo-inositol 4,5-bisphosphate (Gro-PIP2), inhibited PI-PLC I in a concentration-dependent manner. Assays performed with 10 microM [3H]phosphatidylinositol ([3H]PI), 10 microM [3H]phosphatidylinositol 4-phosphate ([3H]PIP) or 10 microM [3H]phosphatidylinositol 4,5-bisphosphate ([3H]PIP2) as substrates, with increasing [Gro-PIP] revealed an IC50 = 380 microM. Kinetic studies with increasing [3H]PI substrate concentrations in the presence of 100 microM and 300 microM Gro-PIP demonstrated that Gro-PIP exhibited competitive inhibition; Kis = 40 microM. Ca2+ concentrations over the range 1.1 microM to 1 mM did not effect inhibition, suggesting that Gro-PIP inhibition of [3H]PI hydrolysis was calcium-independent. To determine whether Gro-PIP was a substrate, 20 microM and 500 microM [3H]Gro-PIP were incubated with PI-PLC I. Anion-exchange HPLC analysis revealed no [3H]IP2 product formation, indicating that [3H]Gro-PIP was not hydrolyzed. Assays performed with [3H]PI and [3H]PIP substrates in the presence of 500 microM [3H]Gro-PIP revealed approx. 75% less [3H]inositol 1-phosphate ([3H]IP1) and [3H]inositol 1,4-bisphosphate ([3H]IP2) product formation, respectively, indicating that [3H]Gro-PIP inhibited the hydrolysis of the substrates by PI-PLC I. These data suggest that Gro-PIP does not serve as a substrate, and that it inhibits PI-PLC I by competitive inhibition in a Ca2(+)-independent fashion.  相似文献   

3.
4.
Appreciable amounts of inositol 1,2-cyclic 4,5-trisphosphate (cIP3) are formed on agonist stimulation of secretory cells, e.g., pancreas (1,2) and parotid (3,4). However, the physiological role of this compound is unknown. To obtain sufficient amounts of cIP3, we have developed a synthetic method to produce cIP3 from inositol 1,4,5-trisphosphate (I(1,4,5)P3). The method is an adaptation of the dicyclohexylcarbodiimide (DCCD) method of Khorana et al. (5), which was originally developed to synthesize 2',3'-cyclic ribonucleotides. The method involves treatment of the pyridinium salt of I(1,4,5)P3 with DCCD in pyridine water, which cyclizes part of the 1-phosphate on the inositol ring to the 1,2-cyclic phosphate. The compound identified as cIP3 cochromatographed with authentic cIP3 in two HPLC systems and on ionophoresis. It was converted to I(1,4,5)P3 on mild acid treatment--a characteristic of cyclic inositol phosphates. Inositol 1,2-cyclic 4,5-trisphosphate is then purified by HPLC. Sufficient amounts of cIP3 can be prepared by this method to carry out numerous experiments on its possible cellular role.  相似文献   

5.
Feng J  Roberts MF  Drin G  Scarlata S 《Biochemistry》2005,44(7):2577-2584
Phosphatidylinositol-specific phospholipase C (PLC) enzymes catalyze the hydrolysis of phosphatidylinositol 4,5 bisphosphate in a two step reaction that involves a cyclic intermediate. The PLCbetafamily are activated by both the alpha and betagamma subunits of heterotrimeric G proteins. To determine which catalytic step is affected by Gbetagamma subunits, we compared the change in PLCbeta(2) activity catalysis toward monomeric short-chain phosphatidylinositol (PI) substrates and monomeric water-soluble cyclic inositol phosphates as well as long-chain PI in bilayer and micellar interfaces in the absence and presence of Gbetagammasubunits. Unlike other PLC enzymes, no cyclic products were detected for either wild-type PLCbeta(2) or a chimeric protein composed of the PH domain of PLCbeta(2) and the catalytic domain of PLCdelta(1). Using cIP as a substrate to examine the second step of the reaction, we found that the presence of Gbetagamma subunits stimulated this step by a higher level than that for the overall reaction (k(cat) 1.5-fold (cIP) as opposed to 1.20-fold for soluble diC(4)PI). Detergents above their CMC can generate the same kinetic activation of PLCbeta(2) as Gbetagamma, suggesting that hydrophobic compounds stabilize the activated state of the enzyme. The most pronounced effect of Gbetagamma is that it relieves competitive product inhibition. Taken together, our results show that activation of PLCbeta(2) occurs through enhancement in the catalytic rate of hydrolysis of the cyclic intermediate and increased product release, and that hydrophobic interactions play a key role.  相似文献   

6.
Inositol phosphate action in an intact cell has been investigated by intracellular microinjection of eight inositol phosphate derivatives into Xenopus laevis oocytes. These cells have calcium-regulated chloride channels but do not have a calcium-induced calcium release system. Microinjection of inositol 1,3,4,5-tetrakisphosphate (IP4), inositol 1,2-(cyclic)-4,5-trisphosphate (cIP3), inositol 1,4,5-trisphosphate (IP3), or inositol 4,5-bisphosphate [(4,5)IP2], open chloride channels to induce a membrane depolarization. However, inositol 1-phosphate (IP1), inositol 1,3,4,5,6-pentakisphosphate (IP5), inositol 1,4-bisphosphate, or inositol 3,4-bisphosphate are unable to induce this depolarization. The depolarization is mimicked by calcium microinjection, inhibited by EGTA coinjection, and is insensitive to removal of extracellular calcium. By means of the depolarization response, the efficacy of various inositol phosphate derivatives are compared. IP3 and cIP3 induce similar half-maximal, biphasic depolarization responses at an intracellular concentration of approximately 90 nM, whereas IP4 induces a mono- or biphasic depolarization at approximately 3400 nM. At concentrations similar to that required for IP3 and cIP3, (4,5)IP2 induces a long-term (greater than 40 min) depolarization. The efficacy (cIP3 = IP3 = (4,5)IP2 much greater than IP4) and action of the various inositol phosphates in an intact cell and their inability to induce meiotic cell division are discussed.  相似文献   

7.
We investigated whether Al(3+)-mediated changes in membrane fluidity can affect the activity of prokaryotic enzymes phospholipase C (PLC) and phospholipase C-phosphatidyl inositol specific (PI-PLC) in liposomes of phosphatidyl choline (PC), PC:phosphatidyl inositol (PI), or PC and polyphosphoinositides (PPI). Al(3+) (10-100 microM) promoted membrane rigidification, evaluated with the probes 1,6-diphenyl-1,3,5-hexatriene and Laurdan, and followed the order: PC:PPI>PC:PI>PC. Al(3+) (25 and 50 microM) did not affect PLC-mediated hydrolysis of PC, PI and PIP(2), but stimulated PIP hydrolysis (48.6%). PI-PLC did not affect PC, PI, and PIP concentrations, but caused a 67% decrease in PIP(2). Al(3+) significantly inhibited PIP(2) hydrolysis in a concentration-dependent (25-50 microM) manner. Results suggest that the inhibition of PIP(2) hydrolysis by Al(3+) could be partially due to a higher lipid packing induced by Al(3+) which could affect the interaction between the enzyme and its substrate.  相似文献   

8.
Aminoacylase has been employed as a model system to study its catalytic properties at low water concentrations/water activities with different water-miscible organic cosolvents. Cosolvents assayed were alcohols and polyols with pure logarithm of the partition coefficient (log P) values, on the standard water/octanol system, ranging between -5.2 and 0.24.

Experimental hydrolysis equilibrium constants (Kapp), at a constant water concentration, decreased with the fall in log P of the cosolvent, as well as with reduction of the water concentration/water activity, as would be expected. The enzyme hydrolytic and synthetic activities, measured at a constant water concentration/water activity value, followed a sigmoidal dependence on log P of the cosolvent employed when the water concentration or water activity values were lower than 50% (w/w) or 0.66, respectively. This became a hyperbolic relationship at higher water concentration/water activity values. A linear relationship between the logarithm of the limiting water activity necessary to maintain enzyme activity and log P was obtained. Both hydrolytic and synthetic activities were suppressed for water activities higher than 0.66 and cosolvents with log P lower than -1.6.  相似文献   

9.
Effects of cryoprotectants on enzyme structure   总被引:2,自引:0,他引:2  
A L Fink 《Cryobiology》1986,23(1):28-37
The interaction between organic cosolvents and proteins is considered, especially from the point of view of effects on protein stability. It is concluded that each protein-cosolvent system constitutes a unique situation, making generalized predictions of expected effects difficult. Two classes of cosolvents are distinguished, based on the nature of their interactions with the protein surface. The thermodynamic instability to the system introduced by the presence of the cosolvent can be accommodated (i) by preferential exclusion of the cosolvent from the vicinity of the protein, (ii) by major structural changes of the protein, or (iii) by aggregation. Polyols tend to undergo preferential exclusion due to unfavorable interactions with nonpolar surface groups, whereas monohydric alcohols and other more hydrophobic cosolvents may undergo preferential exclusion due to adverse interactions with charged groups on the protein surface. Typical cosolvent effects on the structural and catalytic properties of enzymes are illustrated with data for ribonuclease and beta-lactamase with alcohol cosolvents. The relative hydrophobicity of the cosolvent is the major determinant of the effect of a cryosolvent on the enzyme stability and properties. Thus the position of the unfolding transition in cryosolvent will be decreased more by a more nonpolar cosolvent. Different cosolvents can have significantly different effects on the catalytic and structural properties of the same enzyme. Conversely the same cosolvent can have significantly different effects on similar proteins. The number and distribution of the nonpolar and charged groups on the protein's surface probably are the major determinants of the protein contribution to the solvent-protein interaction. The large temperature dependence of the rates of protein unfolding and refolding can be beneficially utilized in cryoprotectant studies of living cells.  相似文献   

10.
Hydrolysis-resistant analogues of GTP specifically stimulate the formation of [3H]inositol mono-, bis- and trisphosphates by saponin-permeabilized Swiss 3T3 cells prelabelled with [3H]inositol. Each inositol phosphate is formed largely by hydrolysis of its parent lipid and not by dephosphorylation of inositol 1,4,5-trisphosphate [(1,4,5)IP3]. Although hydrolysis of phosphatidylinositol 4,5-bisphosphate (PIP2) is most sensitive to guanine nucleotides, hydrolysis of phosphatidyl-inositol (PI) and phosphatidylinositol 4-phosphate (PIP) is quantitatively more important. These results suggest that a guanine nucleotide-dependent regulatory protein(s) (G-protein) is involved in regulating the hydrolysis of PI and PIP, as well as PIP2, and so may allow formation of diacylglycerol (DG) without simultaneous production of (1,4,5)IP3 and mobilization of intracellular Ca2+.  相似文献   

11.
An intrinsic problem often involved in biotransformations carried out by immobilized cells is the poor solubility of substrate and product in water. Increase in hydrophobic substrate availability to such gel-entrapped cells may be attained by the replacement of a fraction of the aqueous medium by water-miscible solvents (cosolvents). The introduction of cosolvents results in increased solubility, but may simultaneously affect enzymic activity and stability. Recently, criteria and guidelines for cosolvent selection on the basis of its effect on intracellular enzyme stability were reported (Freeman, A., and Lilly, M.D. (1987) Appl. Microbiol. Biotechnol. 25, 495-501). In order to understand the impact of the preferable or unsuitable cosolvents on enzyme kinetics and stability, the effects of 1-5 M concentrations of a series of cosolvents (e.g., ethylene glycol, dimethylsulfoxide, N,N-dimethylformamide, ethanol) on a well-characterized, highly specific enzyme model (glucose oxidase) were investigated. The presence of 1-5 M of the cosolvents studied imposed 10-50% reduction in Vmax of the enzyme, but Km was only mildly affected (+/- 25%). This inhibition was attributed to cosolvent effect on small, reversible, conformational changes in the enzyme native structure. Determination of the rate constant of thermal inactivation (at 55 degrees C) of glucose oxidase, in the presence of cosolvents, was employed for the quantitative evaluation of cosolvent effect on enzyme stability. A clear pattern of cosolvent preference in respect to its denaturing effect was obtained, which was identical to the pattern previously observed in a study of oxidoreductases operating from within a whole cell. In both cases diols (e.g., ethylene glycol) were found to be the preferable group of cosolvents. Our results indicate that a soluble enzyme and an intracellular enzyme operating from a whole cell are affected by cosolvents via the same mechanism.  相似文献   

12.
Hydrolytic activity of penicillin V acylase (EC 3.5.1.11) can be improved by using organic cosolvents in monophasic systems. However, the addition of these solvents may result in loss of stability of the enzyme. The thermal stability of penicillin V acylase from Streptomyces lavendulae in water-organic cosolvent monophasic systems depends on the nature of the organic solvent and its concentration in the media. The threshold solvent concentration (at which half enzymatic activity is displayed) is related to the denaturing capacity of the solvent. We found out linear correlations between the free energy of denaturation at 40 degrees C and the concentration of the solvent in the media. On one hand, those solvents with logP values lower than -1.8 have a protective effect that is enhanced when its concentration is increased in the medium. On the other hand, those solvents with logP values higher than -1.8 have a denaturing effect: the higher this value and concentration, the more deleterious. Deactivation constants of PVA at 40 degrees C can be predicted in any monophasic system containing a water-miscible solvent.  相似文献   

13.
Glycosyl phosphoinositol (GPI) anchors on proteins can be modified by palmitoylation of their inositol residue, which makes such anchors resistant to cleavage by phosphatidylinositol-specific phospholipase C (PI-PLC) (Roberts, W. L., Myher, J. J., Kuksis, A., Low, M. G., and Rosenberry, T.L. (1988) J. Biol. Chem. 263, 18766-18775). Mannosylated GPI lipids made in trypanosomal and mammalian cells can also be inositol-acylated, indicating that inositol acylation may be a normal step in GPI anchor synthesis. We find that Saccharomyces cerevisiae mutants blocked in dolichyl phosphate mannose synthesis accumulate a lipid that can be radiolabeled in vivo with [3H]myo-inositol, [3H]GlcN, and [3H]palmitic acid. This lipid is resistant to PI-PLC, yet sensitive to mild alkaline hydrolysis, and has been characterized as GlcN-phosphatidylinositol (PI), fatty acylated on its inositol residue. When yeast membranes are incubated with UDP-[14C] GlcNAc, 14C-labeled GlcNAc-PI and GlcN-PI are made. Addition of ATP and CoA, or of palmitoyl-CoA to incubations results in the synthesis of [14C]GlcN-(acyl-inositol)PI. This lipid is also made when membranes are incubated with [1-14C]palmitoyl-CoA and UDP-GlcNAc. We propose that acyl CoA is the donor in inositol acylation of GlcN-PI, and that GlcN-(acyl-inositol)PI is an obligatory intermediate in GPI synthesis.  相似文献   

14.
The issue as to whether there is direct phosphodiesteratic cleavage of phosphatidylinositol (PI), in addition to that of phosphatidylinositol 4,5-bisphosphate (PIP2), on agonist stimulation of cells has been controversial. In an attempt to resolve this issue, we have studied the kinetics of the formation and breakdown of the cyclic inositol phosphates. This approach is fairly straightforward, since the turnover of the cyclic inositol phosphates is very slow as compared to that of the noncyclic inositol phosphates and proceeds from inositol 1:2-cyclic 4,5-trisphosphate to inositol 1:2-cyclic phosphate (I(c1:2)P) directly by dephosphorylation without any branching pathways, in contrast to the multiple branchpoints of the noncyclic inositol phosphate pathway. Mouse pancreatic minilobules were prelabeled with [3H]inositol for 30 min, followed by washing to remove free inositol. They were then stimulated with carbachol for 30 min. The inositol cyclic polyphosphates reached steady state at 10-15 min, and I(c1:2)P reached steady state at 25 min. We blocked the action of carbachol by addition of an excess of atropine at 30 min, and the rate of disappearance of the three cyclic inositol phosphates was measured. From these data, the contribution of the inositol cyclic polyphosphate pathway to I(c1:2)P was calculated, which was 40-50% of total I(c1:2)P formation. Thus, 40-50% of the I(c1:2)P formed must have been derived from direct phosphodiesteratic cleavage of PI. This approach should prove useful in measuring the relative contributions of PI hydrolysis and PI phosphorylation (phosphatidylinositol 4,5-bisphosphate hydrolysis) in the overall PI cascade.  相似文献   

15.
Penicillin G acylase (PGA) from Kluyvera citrophila immobilized on Amberzyml was used for enantioselective hydrolysis of N-phenylacetylated-dl-tert-leucine (N-Phac-dl-Tle) to produce l-tert-leucine (l-Tle). The effects of various organic cosolvents on hydrolysis of N-Phac-dl-Tle have been investigated in aqueous-cosolvent medium. It was founded that the rate of PGA-catalyzed reaction was significantly affected by the presence of 2% (v/v) organic cosolvent concentration. The initial rate fell with increasing logP of the cosolvent, but for logP values less than −0.24 the rate was faster than in purely aqueous medium. Additionally, the relative rate increases with the increase of dielectric constant (ε) of organic cosolvents. The yields of l-Tle in all aqueous-cosolvent systems were above 95% with the enantiomeric excess (ee) of >99%.  相似文献   

16.
Phosphoinositide-specific phospholipase C (PI-PLC) from human platelet cytosol was purified 190-fold to a specific activity of 0.68 mumol of phosphatidylinositol (PI) cleaved/min per mg of protein. It hydrolyses PI and phosphatidylinositol 4,5-bisphosphate (PIP2), but not phosphatidylcholine, phosphatidylserine or phosphatidylethanolamine. The enzyme exhibits an acid pH optimum of 5.5 and has a molecular mass of 98 kDa as determined by Sephacryl S-200 gel filtration. It required millimolar concentrations of Ca2+ for PI hydrolysis, whereas micromolar concentrations are optimal for PIP2 hydrolysis. Mg2+ could substitute for Ca2+ when PIP2, but not PI, was used as the substrate. EDTA was more effective than EGTA in inhibiting the basal PI-PLC activity towards PIP2. Sodium deoxycholate strongly inhibits the purified PI-PLC activity with either PI or PIP2 as substrate. Ras proteins, either alone or in the form of liposomes, have no effect on PI-PLC activity.  相似文献   

17.
Purified membrane vesicles isolated from sea urchin eggs form nuclear envelopes around sperm nuclei following GTP hydrolysis in the presence of cytosol. A low density subfraction of these vesicles (MV1), highly enriched in phosphatidylinositol (PtdIns), is required for nuclear envelope formation. Membrane fusion of MV1 with a second fraction that contributes most of the nuclear envelope can be initiated without GTP by an exogenous bacterial PtdIns-specific phospholipase C (PI-PLC) which hydrolyzes PtdIns to form diacylglycerides and inositol 1-phosphate. This PI-PLC hydrolyzes a subset of sea urchin membrane vesicle PtdIns into diglycerides enriched in long chain, polyunsaturated species as revealed by a novel liquid chromatography-mass spectrometry analysis. Large unilammelar vesicles (LUVs) enriched in PtdIns can substitute for MV1 in PI-PLC induced nuclear envelope formation. Moreover, MV1 prehydrolyzed with PI-PLC and washed to remove inositols leads to spontaneous nuclear envelope formation with MV2 without further PI-PLC treatment. LUVs enriched in diacylglycerol mimic prehydrolyzed MV1. These results indicate that production of membrane-destabilizing diglycerides in membranes enriched in PtdIns may facilitate membrane fusion in a natural membrane system and suggest that MV1, which binds only to two places on the sperm nucleus, may initiate fusion locally.  相似文献   

18.
Phosphoinositide-specific phospholipase C-delta1 (PI-PLC-delta1) cleaves phosphatidylinositol 4,5-bisphosphate (PI-4,5-P(2), 1), 5-phosphate (PI-5-P, 2) and 4-phosphate (PI-4-P, 3) to form the mixture of the corresponding 4,5-, 5- and 4-phosphorylated inositol 1,2-cyclic phosphate (IcP) and 1-phosphate (IP) (4-6 and 7-9, respectively). In this work, we have studied the rates of the cleavage and the ratios of the cyclic-to-acyclic phosphate products under various pH and Ca(2+) concentration conditions using 31P NMR to monitor the reactions. In agreement with the previous report (Kim et al. Biochim. Biophys. Acta 1989, 163, 177), our results indicate that the IcP/IP ratios strongly depend on the reaction conditions, with the cyclic phosphate products formed predominantly at low pH (pH 5.0) and high calcium concentration (5 mM). Surprisingly, however, we have found that at pH 8.0 and 5 mM Ca(2+), PI-5-P rather than PI-4,5-P(2) is the most preferred substrate with the highest V(max). The cleavage of PI-5-P generated also more cyclic phosphate product than the other two substrates. In addition, we have studied the analogous reaction of phosphorothioate analogues of 1 with the sulfur placed in the nonbridging (10) or bridging (13) positions. We have found that the phosphorothioate analogue 10 produced exclusively the cyclic product 11, whereas the analogue 13 afforded exlusively the acyclic product 7. These results are discussed in terms of the mechanism of PI-PLC, where the cyclic product is formed by 'leaking' from the active site before its subsequent hydrolysis. The potential significance of the cyclic products in the signaling pathways is also discussed.  相似文献   

19.
Enzymes usually undergo rapid inactivation in the presence of organic media. In some cases, the mechanism is quite simple. For example, when an enzyme, fully dispersed and immobilized inside porous supports, is inactivated, at neutral pH and moderate temperature, in the presence of medium-high concentrations of water-miscible organic cosolvents, the unique cause of inactivation is the interaction of the enzyme with cosolvent molecules and the only inactivating effect is the promotion of conformational changes on enzyme structure.

On this basis, two distinct strategies for stabilization of enzymes against organic solvents are proposed:

a. reduction of the causes of inactivation: generation of hyper-hydrophilic micro-environments having a very open structure and fully surrounding every enzyme molecule;

b. reduction of the effects of inactivation: “rigidification of enzymes” via multipoint covalent immobilization.

By using penicillin G acylase (PGA) as a model enzyme, both strategies have been evaluated and compared. Both stabilizing strategies had significant effects. In this case, hydrophilization of the enzyme nano-environment was found to be more effective than rigidification of the enzyme via multipoint covalent attachment. The combined effect of both stabilizing strategies was also tested: multipoint covalently immobilized enzyme molecules were completely surrounded by hyper-hydrophilic microenvironments. In this way, native PGA that was unstable against organic cosolvents (completely inactivated in less than 3 min in 95% dioxane) was transformed into a very stable immobilized derivative (preserving more than 80% of activity after 40 days under the same conditions).  相似文献   

20.
Within 5 min of the binding of anti-mu antibody (anti-mu) to surface IgM on LA350, a human lymphoblastoid B-cell line, there was a significantly enhanced incorporation of 32P into the phosphatidic acid (PA) and phosphatidylinositol (PI) fractions of cellular phospholipids and the magnitude of the early increase in PA was twice as great as that in PI. This anti-mu-enhanced incorporation of 32P into PA and PI required the binding of a divalent form of antibody (IgG or F(ab')2), was blocked by coincubation with micromolar concentrations of soluble IgM, was decreased by incubation of cells at temperatures lower than 37 degrees C, and was inhibited by coincubation with millimolar concentrations of dibutyryl cyclic AMP and theophylline. Similar incorporation studies with [3H]inositol demonstrated a selective and significant increase in labeling of PI. In LA350 labeled with [3H]inositol for 30 hr (equilibrium) and acutely stimulated by anti-mu, specific hydrolysis of phosphorylated PI (PI 4,5-bisphosphate) was measured by the significantly increased release at 15 min of radioactive inositol 1,4,5-trisphosphate, inositol 1,4 bisphosphate, and inositol 1-phosphate. The release of these inositol phosphates was significantly augmented by coincubation with 0.01 M LiCl which prevented their simultaneous enzymatic degradation. All of these findings are consistent with an activation of a linked series of metabolic events known as the PI cycle. In similar cell cultures anti-mu significantly stimulated the secretion of IgM by LA350 as measured at 48 hr in a reverse hemolytic plaque assay. Two other IgM-bearing human lymphoblastoid B-cell lines which gave no evidence of turnover of 32P in PA and PI in response to binding by anti-mu likewise failed to enhance their secretion of IgM. We conclude that the binding of surface IgM on LA350 by anti-mu results in the generation of a transmembrane signal which causes a rapid activation of the PI cycle which itself may play a role in the subsequent increase in IgM secretion.  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