首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The activation of protein kinase C alpha was studied by using a lipid system consisting of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine (POPS) (molar ratio 4:1) and different proportions of 1-palmitoyl-2-oleoyl-sn-glycerol (POG). The phase behavior of the lipidic system was characterized by using differential scanning calorimetry and 31P NMR, and a phase diagram was elaborated. The results suggested the formation of two diacylglycerol/phospholipid complexes, one at 15 mol % of POG and the second at 30 mol % of POG. These two complexes would define the three regions of the phase diagram: in the first region (concentrations of POG lower than 15 mol %) there is gel-gel immiscibility at temperatures below that of the phase transition between C1 and pure phospholipid, and a fluid lamellar phase above of the phase transition. In the second region (between 15 and 30 mol % of POG), gel-gel immiscibility between C1 and C2 with fluid-fluid immiscibility was observed, while inverted hexagonal HII and isotropic phases were detected by 31P NMR. In the third region (concentrations of POG higher than 30 mol %), gel-gel immiscibility seemed to occur between C2 and pure POG along with fluid-fluid immiscibility, while an isotropic phase was detected by 31P NMR. When PKC alpha activity was measured, as a function of POG concentration, maximum activity was found at POG concentrations as low as 5-10 mol %; the activity slightly decreased as POG concentration was increased to 45 mol % at 32 degrees C (above Tc) whereas activity did not change with increasing concentrations of POG at 5 degrees C (below Tc). When the activity was studied as a function of temperature, at different POG concentrations, and depicted as Arrhenius plots, it was found that the activity increased with increasing temperatures, showing a discontinuity at a temperature very close to the phase transition of the system and a lower activation energy at the upper slope of the graph, indicating that the physical state of the membrane affected the interaction of PKC alpha with the membrane.  相似文献   

2.
Lipid activation of protein kinase C alpha (PKC alpha) was studied by using a model mixture containing 1, 2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC), 1, 2-dimyristoyl-sn-glycero-3-phosphoserine (DMPS), and 1, 2-dimyristoyl-sn-glycerol (1,2-DMG). This lipid mixture was physically characterized by differential scanning calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR), and 31P-nuclear magnetic resonance (31P-NMR). Based on these techniques, a phase diagram was constructed by keeping a constant DMPC/DMPS molar ratio of 4:1 and changing the concentration of 1,2-DMG. This phase diagram displayed three regions and two compounds: compound 1 (C1), with 45 mol% 1,2-DMG, and compound 2 (C2), with 60 mol% 1,2-DMG. When the phase diagram was elaborated in the presence of Ca2+ and Mg2+, at concentrations similar to those used in the PKC alpha activity assay, the boundaries between the regions changed slightly and C1 had 35 mol% 1,2-DMG. The activity of PKC alpha was studied at several temperatures and at different concentrations of 1,2-DMG, with a maximum of activity reached at 30 mol% 1,2-DMG and lower values at higher concentrations. In the presence of Ca2+ and Mg2+, maximum PKC alpha activity occurred at concentrations of 1,2-DMG that were close to the boundary in the phase diagram between region 1, where compound C1 and the pure phospholipid coexisted in the gel phase, and region 2, where compounds C1 and C2 coexisted. These results suggest that the membrane structure corresponding to a mixture of 1,2-DMG/phospholipid complex and free phospholipid is better able to support the activity of PKC alpha than the 1,2-DMG/phospholipid complex alone.  相似文献   

3.
The phases and transition sequences for aqueous dispersions of mixtures of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and 1,2-dipalmitoyl-sn-glycerol (1,2-DPG) have been studied by differential scanning calorimetry, dynamic x-ray diffraction, freeze-fracture electron microscopy, 31P-nuclear magnetic resonance spectroscopy, and Fourier-transform infrared spectroscopy. The results have been used to construct a dynamic phase diagram of the binary mixture as a function of temperature over the range 20 degrees-90 degrees C. It is concluded that DPPC and 1,2-DPG form two complexes in the gel phase, the first one with a DPPC/1,2-DPG molar ratio of 55:45 and the second one at a molar ratio of approximately 1:2, defining three different regions in the phase diagram. Two eutectic points are postulated to occur: one at a very low 1,2-DPG concentration and the other at a 1,2-DPG concentration slightly higher than 66 mol%. At temperatures higher than the transition temperature, lamellar phases were predominant at low 1,2-DPG concentrations, but nonlamellar phases were found to be predominant at high proportions of 1,2-DPG. A very important aspect of these DPPC/1,2-DPG mixtures was that, in the gel phase, they showed a ripple structure, as seen by freeze-fracture electron microscopy and consistent with the high lamellar repeat spacings seen by x-ray diffraction. Ripple phase characteristics were also found in the fluid lamellar phases occurring at concentrations up to 35.6 mol% of 1,2-DPG. Evidence was obtained by Fourier transform infrared spectroscopy of the dehydration of the lipid-water interface induced by the presence of 1,2-DPG. The biological significance of the presence of diacylglycerol in membrane lipid domains is discussed.  相似文献   

4.
PMP1 is a 38-residue plasma membrane protein of the yeast Saccharomyces cerevisiae that regulates the activity of the H(+)-ATPase. The cytoplasmic domain conformation results in a specific interfacial distribution of five basic side chains, thought to strongly interact with anionic phospholipids. We have used the PMP1 18-38 fragment to carry out a deuterium nuclear magnetic resonance ((2)H-NMR) study for investigating the interactions between the PMP1 cytoplasmic domain and phosphatidylserines. For this purpose, mixed bilayers of 1-palmitoyl, 2-oleoyl-sn-glycero-3-phosphocholine (POPC) and 1-palmitoyl, 2-oleoyl-sn-glycero-3-phosphoserine (POPS) were used as model membranes (POPC/POPS 5:1, m/m). Spectra of headgroup- and chain-deuterated POPC and POPS phospholipids, POPC-d4, POPC-d31, POPS-d3, and POPS-d31, were recorded at different temperatures and for various concentrations of the PMP1 fragment. Data obtained from POPS deuterons revealed the formation of specific peptide-POPS complexes giving rise to a slow exchange between free and bound PS lipids, scarcely observed in solid-state NMR studies of lipid-peptide/protein interactions. The stoichiometry of the complex (8 POPS per peptide) was determined and its significance is discussed. The data obtained with headgroup-deuterated POPC were rationalized with a model that integrates the electrostatic perturbation induced by the cationic peptide on the negatively charged membrane interface, and a "spacer" effect due to the intercalation of POPS/PMP1f complexes between choline headgroups.  相似文献   

5.
The phase behaviour of 1-palmitoyl-2-oleoylphosphatidylethanolamine (POPE) was studied by differential scanning calorimetry and 31P-NMR spectroscopy. Modulation of the phase behaviour of POPE by 1-palmitoyl-2-oleoylphosphatidylserine (POPS). 1-palmitoyl-2-oleoylphosphatidylcholine (POPC), 1,2-di-olein (DOG), CaCl2, MgCl2, and combinations of these substances was studied. The bilayer-forming lipids, POPS and POPC, raise the bilayer-to-hexagonal phase-transition temperature of POPE. The POPC has a greater effect than POPS, probably because the former lipid is more miscible with POPE. Addition of 10 mM CaCl2 has little effect on the phase-transitions of POPE/POPC mixtures, but it greatly decreases the effectiveness of POPS in raising the bilayer-to-hexagonal phase-transition temperature of POPE. The effectiveness of DOG in lowering the phase-transition temperature of POPE is also greatly reduced in the presence of 10 mM CaCl2. This phenomenon may play a role in the negative feedback regulation of protein kinase C.  相似文献   

6.
The protein kinase C (PKC) isoforms are maintained in an inactive and closed conformation by intramolecular interactions. Upon activation these are disrupted by activators, binding proteins and cellular membrane. We have seen that autophosphorylation of two sites in the C-terminal V5 domain is crucial to keep PKC alpha insensitive to the activator diacylglycerol, which presumably is caused by a masking of the diacylglycerol-binding C1a domain. Here we demonstrate that the diacylglycerol sensitivity of the PKC beta isoforms also is suppressed by autophosphorylation of the V5 sites. To analyze conformational differences, a fusion protein ECFP-PKC alpha-EYFP was expressed in cells and the FRET signal was analyzed. The analogous mutant with autophosphorylation sites exchanged for alanine gave rise to a substantially lower FRET signal than wild-type PKC alpha indicating a conformational difference elicited by the mutations. Expression of the isolated PKC alpha V5 domain led to increased diacylglycerol sensitivity of PKC alpha. We identified acidic residues in the V5 domain that, when mutated to alanines or lysines, rendered PKC alpha sensitive to diacylglycerol. Furthermore, mutation to glutamate of four lysines in a lysine-rich cluster in the C2 domain gave a similar effect. Simultaneous reversal of the charges of the acidic residues in the V5 and the lysines in the C2 domain gave rise to a PKC alpha that was insensitive to diacylglycerol. We propose that these structures participate in an intramolecular interaction that maintains PKC alpha in a closed conformation. The disruption of this interaction leads to an unmasking of the C1a domain and thereby increased diacylglycerol sensitivity of PKC alpha.  相似文献   

7.
The behavior of cholesterol is different in mixtures with phosphatidylcholine as compared with phosphatidylserine. In (13)C cross polarization/magic angle spinning nuclear magnetic resonance spectra, resonance peaks of the vinylic carbons of cholesterol are a doublet in samples containing 0.3 or 0.5 mol fraction cholesterol with 1-palmitoyl-2-oleoyl phosphatidylserine (POPS) or in cholesterol monohydrate crystals, but a singlet with mixtures of cholesterol and 1-palmitoyl-2-oleoyl phosphatidylcholine (POPC). At these molar fractions of cholesterol with POPS, resonances of the C-18 of cholesterol appear at the same chemical shifts as in pure cholesterol monohydrate crystals. These resonances do not appear in samples of POPS with 0.2 mol fraction cholesterol or with POPC up to 0.5 mol fraction cholesterol. In addition, there is another resonance from the cholesterol C18 that appears in all of the mixtures of phospholipid and cholesterol but not in pure cholesterol monohydrate crystals. Using direct polarization, the fraction of cholesterol present as crystallites in POPS with 0.5 mol fraction cholesterol is found to be 80%, whereas with the same mol fraction of cholesterol and POPC none of the cholesterol is crystalline. After many hours of incubation, cholesterol monohydrate crystals in POPS undergo a change that results in an increase in the intensity of certain resonances of cholesterol monohydrate in (13)C cross polarization/magic angle spinning nuclear magnetic resonance, indicating a rigidification of the C and D rings of cholesterol but not other regions of the molecule.  相似文献   

8.
Protein kinase Cα (PKCα) is activated by its translocation to the membrane. Activity assays show the importance of PIP(2) in determining the specific activity of this enzyme. A FRET stopped flow fluorescence study was carried out to monitor the rapid kinetics of protein binding to model membranes containing POPC/POPS/DOG and eventually PIP(2). The results best fitted a binding mechanism in which protein bound to the membrane following a two-phase mechanism with a first bimolecular reaction followed by a slow unimolecular reaction. In the absence of PIP(2), the rapid protein binding rate was especially dependent on POPS concentration. Formation of the slow high affinity complex during the second phase seems to involve specific interactions with POPS and DOG since it is only sensitive to changes within relatively low concentration ranges of these lipids. Both the association and dissociation rate constants fell in the presence of PIP(2). We propose a model in which PKCα binds to the membranes via a two-step mechanism consisting of the rapid membrane initial recruitment of PKCα driven by interactions with POPS and/or PIP(2) although interactions with DOG are involved too. PKCα searches on the lipid bilayer in two dimensions to establish interactions with its specific ligands.  相似文献   

9.
The binding of calcium to headgroup deuterated 1-palmitoyl, 2-oleoyl-sn-glycero-3-phosphoserine (POPS) was investigated by using deuterium magnetic resonance in pure POPS membranes and in mixed 1-palmitoyl, 2-oleoyl-sn-glycero-3-phosphocholine (POPC)/POPS 5:1 (m:m) bilayers. Addition of CaCl2 to pure POPS bilayers led to two component spectra attributed, respectively, to liquid-crystallin POPS (less than 15 kHz) and POPS molecules in the calcium-induced dehydrated phase (cochleate) (approximately 120 kHz). The liquid-crystalline component has nearly disappeared at a Ca2+ to POPS ratio of 0.5, indicating that, under such conditions, most of the POPS molecules are in the precipitated cochleate phase. After dilution of the POPS molecules in zwitterionic POPC membranes (POPC/POPS 5:1 m:m), single component spectra characteristic of POPS in the liquid-crystalline state were observed in the presence of Molar concentrations of calcium ions (Ca2+ to POPS ratio greater than 50), showing that the amount of dehydrated cochleate PS-Ca2+ phase, if any, was low (less than 5%) under such conditions. Deuterium NMR data obtained in the 15-50 degrees C temperature range with the mixed PC/PS membranes, either in the absence or the presence of Ca2+ ions, indicate that the serine headgroup undergoes a temperature-induced conformational change, independent of the presence of Ca2+. This is discussed in relation to other headgroup perturbations such as that observed upon change of the membrane surface charge density.  相似文献   

10.
Whereas retinoic acids control nuclear events, a second class of retinol metabolites, that is, the hydroxylated forms exemplified by 14-hydroxy-retro-retinol (HRR), operate primarily in the cytoplasm. They function as regulatory cofactors for cell survival/cell death decisions. In accordance with these biological aspects, we demonstrate that these retinoids bound protein kinase C (PKC) alpha with nanomolar affinity and markedly enhance the activation of PKC alpha and the entire downstream MAP kinase pathway by reactive oxygen species. HRR was 10 times more efficient than retinol, and the optimum doses are 10-7 and 10-6 M, respectively. PKC alpha activation was reversed rapidly by imposition of reducing conditions. The retinoid binding site was mapped to the first cysteine-rich region in the regulatory domain, C1A, yet was distinct from the binding sites of diacylglycerol and phorbol esters. The C1B domain bound retinoids poorly. The emerging theme is that retinoids serve as redox regulators of protein kinase C.  相似文献   

11.
Cationic amphiphiles used for transfection can be incorporated into biological membranes. By differential scanning calorimetry (DSC), cholesterol solubilization in phospholipid membranes, in the absence and presence of cationic amphiphiles, was determined. Two different systems were studied: 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)+cholesterol (1:3, POPC:Chol, molar ratio) and 1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-l-serine] (POPS)+cholesterol (3:2, POPS:Chol, molar ratio), which contain cholesterol in crystallite form. For the zwitterionic lipid POPC, cationic amphiphiles were tested, up to 7 mol%, while for anionic POPS bilayers, which possibly incorporate more positive amphiphiles, the fractions used were higher, up to 23 mol%. 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) and DOTAP in methyl sulfate salt form (DOTAPmss) were found to cause a small decrease on the enthalpy of the cholesterol transition of pure cholesterol aggregates, possibly indicating a slight increase on the cholesterol solubilization in POPC vesicles. With the anionic system POPS:Chol, the cationic amphiphiles dramatically change the cholesterol crystal thermal transition, indicating significant changes in the cholesterol aggregates. For structural studies, phospholipids spin labeled at the 5th or 16th carbon atoms were incorporated. In POPC, at the bilayer core, the cationic amphiphiles significantly increase the bilayer packing, decreasing the membrane polarity, with the cholesterol derivative 3 beta-[N-(N',N'-dimethylaminoethane)-carbamoyl]-cholesterol (DC-chol) displaying a stronger effect. In POPS and POPS:Chol, DC-chol was also found to considerably increase the bilayer packing. Hence, exogenous cationic amphiphiles used to deliver nucleic acids to cells can change the bilayer packing of biological membranes and alter the structure of cholesterol crystals, which are believed to be the precursors to atherosclerotic lesions.  相似文献   

12.
The cationic amphiphile, cholesteryl-3-carboxyamidoethylene-trimethylammonium iodide, can alter the substrate specificity of protein kinase C (PKC). The phosphorylation of histone catalyzed by PKC requires the binding of the enzyme to phospholipid vesicles. This cationic amphiphile reduces both the binding of PKC to lipid and as a consequence its rate of phosphorylation of histone. In contrast, PKC bound to large unilamellar vesicles (LUVs) composed of 50 mol % POPS, 20 mol % POPC, and 30 mol % of this amphiphile catalyzes protamine sulfate phosphorylation by an almost 4 fold greater rate. This activation requires phosphatidylserine (PS) and is inhibited by Ca2+. The extent of activation is affected by the time of incubation of PKC with LUVs. This data suggests a novel mechanism by which PKC-dependent signal transduction pathways may be altered by altering the protein targets of this enzyme.  相似文献   

13.
In this study, the role of interdomain interactions involving the C1 and C2 domains in the mechanism of activation of PKC was investigated. Using an in vitro assay containing only purified recombinant proteins and the phorbol ester, 4 beta-12-O-tetradecanoylphorbol-13-acetate (TPA), but lacking lipids, it was found that PKC alpha bound specifically, and with high affinity, to a alpha C1A-C1B fusion protein of the same isozyme. The alpha C1A-C1B domain also potently activated the isozyme in a phorbol ester- and diacylglycerol-dependent manner. The level of this activity was comparable with that resulting from membrane association induced under maximally activating conditions. Furthermore, it was found that alpha C1A-C1B bound to a peptide containing the C2 domain of PKC alpha. The alpha C1A-C1B domain also activated conventional PKC beta I, -beta II, and -gamma isoforms, but not novel PKC delta or -epsilon. PKC delta and -epsilon were each activated by their own C1 domains, whereas PKC alpha, -beta I, -beta II, or -gamma activities were unaffected by the C1 domain of PKC delta and only slightly activated by that of PKC epsilon. PKC zeta activity was unaffected by its own C1 domain and those of the other PKC isozymes. Based on these findings, it is proposed that the activating conformational change in PKC alpha results from the dissociation of intra-molecular interactions between the alpha C1A-C1B domain and the C2 domain. Furthermore, it is shown that PKC alpha forms dimers via inter-molecular interactions between the C1 and C2 domains of two neighboring molecules. These mechanisms may also apply for the activation of the other conventional and novel PKC isozymes.  相似文献   

14.
C1 domains mediate the recognition and subsequent signaling response to diacylglycerol and phorbol esters by protein kinase C (PKC) and by several other families of signal-transducing proteins such as the chimerins or RasGRP. MRCK (myotonic dystrophy kinase-related Cdc42 binding kinase), a member of the dystrophia myotonica protein kinase family that functions downstream of Cdc42, contains a C1 domain with substantial homology to that of the diacylglycerol/phorbol ester-responsive C1 domains and has been reported to bind phorbol ester. We have characterized here the interaction of the C1 domains of the two MRCK isoforms alpha and beta with phorbol ester. The MRCK C1 domains bind [20-(3)H]phorbol 12,13-dibutyrate with K(d) values of 10 and 17 nm, respectively, reflecting 60-90-fold weaker affinity compared with the protein kinase C delta C1b domain. In contrast to binding by the C1b domain of PKCdelta, the binding by the C1 domains of MRCK alpha and beta was fully dependent on the presence of phosphatidylserine. Comparison of ligand binding selectivity showed resemblance to that by the C1b domain of PKCalpha and marked contrast to that of the C1b domain of PKCdelta. In intact cells, as in the binding assays, the MRCK C1 domains required 50-100-fold higher concentrations of phorbol ester for induction of membrane translocation. We conclude that additional structural elements within the MRCK structure are necessary if the C1 domains of MRCK are to respond to phorbol ester at concentrations comparable with those that modulate PKC.  相似文献   

15.
Electrophysiological recordings from freshly dissociated smooth muscle cells from the stomach of the toad Bufo marinus revealed two types of Ca2+ currents. One has a low threshold of activation and inactivates rapidly; the other has a high threshold of activation and inactivates more slowly. Acetylcholine (ACh) increased the high-threshold current but not the low-threshold current. The synthetic diacylglycerol analog sn-1,2-dioctanoylglycerol, an activator of protein kinase C (PKC), mimicked these effects of ACh on Ca2+ currents. However, another diacylglycerol analog, 1,2-dioctanoyl-3-thioglycerol, which has a closely related structure but does not activate PKC, failed to increase the Ca2+ current. The same was true of 1,2-dioctanoyl-3-chloropropanediol, an analog that even at high concentrations only minimally activates PKC. These results suggest that diacylglycerol may be the second messenger mediating the effects of ACh on one type of voltage-activated Ca2+ channel, possibly by activating PKC.  相似文献   

16.
We have used attenuated total internal reflection infrared spectroscopy (ATR-IR) spectroscopy to study the association of the C2 domain from protein kinase Cα (PKCα) with different phospholipid membranes, so as to characterise the mode of membrane docking and its modulation by the second-messenger lipid PIP?. In parallel, we have also examined the membrane interaction of the C2 domain from cytosolic phospholipase A?. PIP? did not induce significant changes in secondary structure of the membrane-bound PKCα-C2 domain, nor did binding of the PKCα-C2 domain change the dichroic ratios of the lipid chains, whereas the C2 domain from phospholipase A? did perturb the lipid chain orientation. Measurements of the dichroic ratios for the amide I and amide II protein bands were combined so as to distinguish the tilt of the β-sheets from that of the β-strands within the sheet. When associated with POPC/POPS membranes, the β-sandwich of the PKCα-C2 domain is inclined at an angle α=35° to the membrane normal, i.e., is oriented more nearly perpendicular than parallel to the membrane. In the process of membrane docking, the tilt angle increases to α=44° in the presence of PIP?, indicating that the β-sandwich comes closer to the membrane surface, so confirming the importance of this lipid in determining docking of the C2 domain and consequent activation of PKCα.  相似文献   

17.
The interaction of Saposin C (Sap C) with negatively charged phospholipids such as phosphatidylserine (PS) is essential for its biological function. In this study, Sap C (initially protonated in a weak acid) was inserted into multilamellar vesicles (MLVs) consisting of either 1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-L-serine] (negatively charged, POPS) or 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (neutrally charged, POPC). The MLVs were then investigated using solid-state NMR spectroscopy under neutral pH (7.0) conditions. The (2)H and (31)P solid-state NMR spectroscopic data of Sap C-POPS and Sap C-POPC MLVs (prepared under the same conditions) were compared using the (2)H order parameter profiles of the POPC-d(31) or POPS-d(31) acyl chains as well as the (31)P chemical shift anisotropy width and (31)P T(1) relaxation times of the phospholipids headgroups. All those solid-state NMR spectroscopic approaches indicate that protonated Sap C disturbs the POPS bilayers and not the POPC lipid bilayers. These observations suggest for the first time that protonated Sap C inserts into PS bilayers and forms a stable complex with the lipids even after resuspension under neutral buffer conditions. Additionally, (31)P solid-state NMR spectroscopic studies of mechanically oriented phospholipids on glass plates were conducted and perturbation effect of Sap C on both POPS and POPC bilayers was compared. Unlike POPC bilayers, the data indicates that protonated Sap C (initially protonated in a weak acid) was unable to produce well-oriented POPS bilayers on glass plates at neutral pH. Conversely, unprotonated Sap C (initially dissolved in a neutral buffer) did not interact significantly with POPS phospholipids allowing them to produce well-oriented bilayers at neutral pH.  相似文献   

18.
Protein kinase C (PKC) I (gamma), II (beta) and III (alpha) subspecies are all activated by 1,1-di-(p-hydroxyphenyl)ethylene derivatives (DPE) at micromolar concentrations. This PKC activation depends on the presence of both Ca2+ and phosphatidylserine (PS) but does not require diacylglycerol (DG). DPEs enhance PKC activity at low PS concentrations, but not at saturating PS concentrations. Like DG, DPEs increase the apparent affinity of PKC for PS as well as for Ca2+, but lead to a decrease in the catalytic activity (Vmax). In the presence of saturating DG concentrations, DPEs exhibit an inhibitory action. The derivatives also inhibit the activity of the proteolytic fragment of PKC, protein kinase M. It is concluded that DPEs are mixed-type inhibitors, probably interacting with the catalytic domain of the enzyme.  相似文献   

19.
Cationic amphiphiles used for transfection can be incorporated into biological membranes. By differential scanning calorimetry (DSC), cholesterol solubilization in phospholipid membranes, in the absence and presence of cationic amphiphiles, was determined. Two different systems were studied: 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) + cholesterol (1:3, POPC:Chol, molar ratio) and 1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-l-serine] (POPS) + cholesterol (3:2, POPS:Chol, molar ratio), which contain cholesterol in crystallite form. For the zwitterionic lipid POPC, cationic amphiphiles were tested, up to 7 mol%, while for anionic POPS bilayers, which possibly incorporate more positive amphiphiles, the fractions used were higher, up to 23 mol%. 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) and DOTAP in methyl sulfate salt form (DOTAPmss) were found to cause a small decrease on the enthalpy of the cholesterol transition of pure cholesterol aggregates, possibly indicating a slight increase on the cholesterol solubilization in POPC vesicles. With the anionic system POPS:Chol, the cationic amphiphiles dramatically change the cholesterol crystal thermal transition, indicating significant changes in the cholesterol aggregates. For structural studies, phospholipids spin labeled at the 5th or 16th carbon atoms were incorporated. In POPC, at the bilayer core, the cationic amphiphiles significantly increase the bilayer packing, decreasing the membrane polarity, with the cholesterol derivative 3β-[N-(N′,N′-dimethylaminoethane)-carbamoyl]-cholesterol (DC-chol) displaying a stronger effect. In POPS and POPS:Chol, DC-chol was also found to considerably increase the bilayer packing. Hence, exogenous cationic amphiphiles used to deliver nucleic acids to cells can change the bilayer packing of biological membranes and alter the structure of cholesterol crystals, which are believed to be the precursors to atherosclerotic lesions.  相似文献   

20.
Members of the protein kinase C (PKC) family are characterized by an NH2-terminal regulatory domain containing binding sites for calcium, phosphatidylserine, and diacylglycerol (or tumor-promoting phorbol esters), a small central hinge region and a COOH-terminal catalytic domain. We have constructed fusion proteins in which the regulatory domain of PKC alpha was removed and replaced by a 19-amino acid leader sequence containing a myristoylation consensus or by the same sequence in which the amino-terminal glycine was changed to alanine to prevent myristoylation. The goal was to generate constitutively active mutants of PKC that were either membrane bound, due to their myristoylation, or cytoplasmic. Western blotting of fractions from COS cells transfected with plasmids encoding wild-type and mutant proteins revealed that PKC alpha resided entirely in a Triton X-100 soluble (TS) fraction, whereas both the myristoylated and nonmyristoylated mutants were associated primarily with the nuclear envelope fraction. A similar mutant that lacked the 19 amino acid leader sequence was also found almost entirely in the nuclear envelope, as was a truncation mutant containing only the regulatory domain, hinge region, and a small portion of the catalytic domain. However, an additional truncation mutant consisting of only the regulatory domain plus the first one-third of the hinge region was almost entirely in the TS fraction. A nonmyristoylated fusion protein containing only the catalytic domain was also found in the nuclear envelope. Immunostaining of cells transfected with these constructs revealed that both the myristoylated and nonmyristoylated mutants were localized in nuclei, whereas wild-type PKC alpha was primarily cytoplasmic and perinuclear. Phorbol dibutyrate treatment of PKC alpha-transfected cells resulted in increased perinuclear and nuclear staining. The results are consistent with a model in which activation of PKC, by phorbol esters or by deletion of the regulatory domain, exposes regions in the hinge and catalytic domains that interact with a PKC "receptor" present in the nuclear envelope, and may explain the ability of wild-type PKC to be translocated to the nucleus under certain conditions.  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