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 46 毫秒
1.
Interactions between the nicotinic acetylcholine receptor (nAChR) and phosphatidic acid (PA) are bidirectional in that membranes containing PA are effective at stabilizing an agonist-responsive nAChR, whereas incorporation of the nAChR into the same membranes leads to a substantial increase in lipid lateral packing density. A previous study suggested that the ability of PA to adopt a dianionic ionization state is key. We monitored the ionization state of PA in both reconstituted and protein-free membranes. In model membranes composed of PA and 3:2 (mol/mol) phosphatidylcholine (PC)/PA, the monoanionic-to-dianionic transition of PA was detected with a pKa of 8.7 and 6.5, respectively. In the reconstituted 3:2 PC/PA membranes, however, PA was stabilized in a monoanionic state at pH values up to 10. Although dianionic PA does not play a role in nAChR function, we found that both the stabilization of monoanionic PA and the concentration of other cations at the bilayer surface can account for changes in bilayer physical properties that are observed upon incorporation of the nAChR into 3:2 PC/PA membranes. A nAChR-induced concentration of cations at the bilayer surface likely mediates interactions between the nAChR and the anionic lipids in its membrane environment.  相似文献   

2.
The effects of cholesterol (Chol) and an anionic lipid, dioleoylphosphatidic acid (DOPA) on the conformational equilibria of the nicotinic acetylcholine receptor (nAChR) have been investigated using Fourier transform infrared difference spectroscopy. The difference between spectra recorded in the presence and absence of agonist from the nAChR reconstituted into 3:1:1 egg phosphatidylcholine (EPC)/DOPA/Chol membranes exhibits positive and negative bands that serve as markers of the structural changes associated with the resting to desensitized conformational change. These markers are absent in similar difference spectra recorded from the nAChR reconstituted into EPC membranes lacking both Chol and DOPA, indicating that the nAChR cannot undergo conformational change in response to agonist binding. When low levels of either Chol or DOPA up to 25 mol % of the total lipid are included in the EPC membranes, the markers suggest the predominant stabilization of a conformation that is a structural intermediate between the resting and desensitized states. At higher levels of either Chol or DOPA, the nAChR is stabilized in a conformation that is capable of undergoing agonist-induced desensitization, although DOPA appears to be required for the nAChR to adopt a conformation fully equivalent to that found in native and 3:1:1 EPC/DOPA/Chol membranes. The ability of these two structurally diverse lipids, as well as others (Ryan, S. E., Demers, C. N., Chew, J. P., Baenziger, J. E. (1996) J. Biol. Chem. 271, 24590-24597), to modulate the functional state of the nAChR suggests that lipids act on the nAChR via an indirect effect on some physical property of the lipid bilayer. The data also suggest that anionic lipids are essential to stabilize a fully functional nAChR. We propose that membrane fluidity modulates the relative populations of nAChRs in the resting and desensitized states but that subtle structural changes in the presence of anionic lipids are essential for full activity.  相似文献   

3.
Bilayers containing phosphatidylcholine (PC) and the anionic lipid phosphatidic acid (PA) are particularly effective at stabilizing the nicotinic acetylcholine receptor (nAChR) in a functional conformation that undergoes agonist-induced conformational change. The physical properties of PC membranes containing PA are also substantially altered upon incorporation of the nAChR. To test whether or not the negative charge of PA is responsible for this "bi-directional coupling," the nAChR was reconstituted into membranes composed of PC with varying levels of the net negatively charged lipid phosphatidylserine (PS). In contrast to PA, increasing levels of PS in PC membranes do not stabilize an increasing proportion of nAChRs in a functional resting conformation, nor do they slow nAChR peptide hydrogen exchange kinetics. Incorporation of the nAChR had little effect on the physical properties of the PC/PS membranes, as monitored by the gel-to-liquid crystal phase transition temperatures of the bilayers. These results show that a net negative charge alone is not sufficient to account for the unique interactions that occur between the nAChR and PC/PA membranes. Incorporation of the receptor into PC/PS membranes, however, did lead to an altered head group conformation of PS possibly by recruiting divalent cations to the membrane surface. The results show that the nAChR has complex and unique interactions with both PA and PS. The interactions between the nAChR and PS may be bridged by divalent cations, such as calcium.  相似文献   

4.
Dickey AN  Faller R 《Biophysical journal》2008,95(12):5637-5647
It has been found experimentally that negatively charged phosphatidic acid (PA) lipids and cholesterol molecules stabilize the nicotinic acetylcholine receptor (nAChR) in a functional resting state that can participate in an agonist-induced conformational change. In this study, we compare phosphatidylcholine (PC) and PA lipid behavior in the presence of the nAChR to determine why PC lipids do not support a functional nAChR. For lipids that are located within 1.0 nm of the protein, both PC and PA lipids have very similar order-parameter and bilayer-thickness values, which indicate that the annular lipid properties are protein-dependent. The most significant difference between the PC and PA bilayers is the formation of a lipid domain around the protein, which is visible in the PA bilayer but not the PC bilayer. This suggests that the PA domain may help stabilize the nAChR resting state. The PA lipids in the microdomain have a decreased order compared to a homogeneous PA bilayer and the lipids near the protein attempt to increase the free space in their vicinity by residing in multiple lateral planes.  相似文献   

5.
Functional membranes containing purified Torpedo californica acetylcholine receptor and dioleoylphosphatidylcholine (DOPC) were prepared by a cholate dialysis procedure with lipid to protein ratios of 100-400 to 1 (mol/mol). Spin-labeled lipids were incorporated into the reconstituted membranes and into native membranes prepared from Torpedo electroplax, and electron paramagnetic resonance (EPR) spectra were recorded between 0 and 20 degrees C. The spin-labels included nitroxide derivatives of stearic acid (16-doxylstearic acid), androstane, phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylserine (PS), and phosphatidic acid (PA). The phospholipid spin-labels had 16-doxylstearic acid in the sn-2 position. All the spectra showed two components corresponding to a relatively mobile bilayer component and a motionally restricted "protein-perturbed" component. The relative amounts of mobile and perturbed components were quantitated by spectral subtraction and integration techniques. The mobile/perturbed ratio was somewhat temperature dependent, and the results are discussed in terms of exchange between mobile and perturbed environments. Plots of the mobile/perturbed ratios vs. lipid/protein ratios at 1 degree C gave straight lines from which the relative binding affinity of each spin-label and the number of perturbed lipids per receptor protein could be calculated. All the spin-labels gave similar values for the number of perturbed lipids (40 +/- 7), a number close to the number of lipids that will fit around the intramembranous perimeter of the receptor. The affinities of the spin-labeled lipids for the receptor relative to DOPC were androstane (K = 4.3) congruent to 16-doxylstearic acid (4.1) greater than PA (2.7) greater than PE (1.1) approximately PC (1.0) approximately PS (0.7). The lipids having the highest affinity for the acetylcholine receptor were also those that have the largest effects on the ion flux functional properties of the receptor, and the results are discussed in terms of lipid effects on receptor function.  相似文献   

6.
Over the past three decades, the Torpedo californica nicotinic acetylcholine receptor (nAChR) has been one of the most extensively studied membrane protein systems. However, the effects of detergent solubilization on nAChR stability and function are poorly understood. The use of lipid-analog detergents for nAChR solubilization has been shown to preserve receptor stability and functionality. The present study used lipid-analog detergents from phospholipid-analog and cholesterol-analog detergent families for solubilization and affinity purification of the receptor and probed nAChR ion channel function using planar lipid bilayers (PLBs) and stability using analytical size exclusion chromatography (A-SEC) in the detergent-solubilized state. We also examined receptor mobility on the lipidic cubic phase (LCP) by measuring the nAChR mobile fraction and diffusion coefficient through fluorescence recovery after photobleaching (FRAP) experiments using lipid-analog and non-lipid-analog detergents. Our results show that it is possible to isolate stable and functional nAChRs using lipid-analog detergents, with characteristic ion channel currents in PLBs and minimal aggregation as observed in A-SEC. Furthermore, fractional mobility and diffusion coefficient values observed in FRAP experiments were similar to the values observed for these parameters in the recently LCP-crystallized β(2)-adrenergic receptor. The overall results show that phospholipid-analog detergents with 16 carbon acyl-chains support nAChR stability, functionality and LCP mobility.  相似文献   

7.
Interactions of benzophenone (BP) with the Torpedo nicotinic acetylcholine receptor (nAChR) were characterized by electrophysiological analyses, radioligand binding assays, and photolabeling of nAChR-rich membranes with [3H]BP to identify the amino acids contributing to its binding sites. BP acted as a low potency noncompetitive antagonist, reversibly inhibiting the ACh responses of nAChRs expressed in Xenopus oocytes (IC50 = 600 microM) and the binding of the noncompetitive antagonist [3H]tetracaine to nAChR-rich membranes (IC50 = 150 microM). UV irradiation at 365 nm resulted in covalent incorporation of [3H]BP into the nAChR subunits (delta > alpha approximately beta > gamma), with photoincorporation limited to the nAChR transmembrane domain. Comparison of nAChR photolabeling in the closed state (absence of agonist) and desensitized state (equilibrated with agonist) revealed selective desensitized state labeling in the delta subunit of deltaPhe-232 in deltaM1 and deltaPro-286/deltaIle-288 near the beginning of deltaM3 that are within a pocket at the interface between the transmembrane and extracellular domains. There was labeling in the closed state within the ion channel at position M2-13 (alphaVal-255, betaVal-261, and deltaVal-269) that was reduced by 90% upon desensitization and labeling in the transmembrane M3 helices of the beta and gamma subunits (betaMet-285, betaMet-288, and gammaMet-291) that was reduced by 50-80% in the desensitized state. Labeling at the lipid interface (alphaMet-415 in alphaM4) was unaffected by agonist. These results provide a further definition of the regions in the nAChR transmembrane domain that differ in structure between the closed and desensitized states.  相似文献   

8.
Radioligand binding, photoaffinity labeling, and docking and molecular dynamics were used to characterize the tricyclic antidepressant (TCA) binding sites in the nicotinic acetylcholine receptor (nAChR). Competition experiments indicate that the noncompetitive antagonist phencyclidine (PCP) inhibits [3H]imipramine binding to resting (closed) and desensitized nAChRs. [3H]2-azidoimipramine photoincorporates into each subunit from the desensitized nAChR with approximately 25% of the labeling specifically inhibited by TCP (a PCP analog), whereas no TCP-inhibitable labeling was observed in the resting (closed) state. For the desensitized nAChR and within the alpha subunit, the majority of specific [3H]2-azidoimipramine labeling mapped to a approximately 20 kDa Staphylococcus aureus V8 protease fragment (alphaV8-20; Ser173-Glu338). To further map the labeling site, the alphaV8-20 fragment was further digested with endoproteinase Lys-C and resolved by Tricine SDS-PAGE. The principal labeled fragment (11 kDa) was further purified by rpHPLC and subjected to N-terminal sequencing. Based on the amino terminus (alphaMet243) and apparent molecular weight, the 11 kDa fragment contains the channel lining M2 segment. Finally, docking and molecular dynamics results indicate that imipramine and PCP interact preferably with the M2 transmembrane segments in the middle of the ion channel. Collectively, these results are consistent with a model where PCP and TCA bind to overlapping sites within the lumen of the Torpedo nAChR ion channel.  相似文献   

9.
Chiara DC  Dangott LJ  Eckenhoff RG  Cohen JB 《Biochemistry》2003,42(46):13457-13467
To identify inhalational anesthetic binding domains in a ligand-gated ion channel, we photolabeled nicotinic acetylcholine receptor (nAChR)-rich membranes from Torpedo electric organ with [(14)C]halothane and determined by Edman degradation some of the photolabeled amino acids in nAChR subunit fragments isolated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and high-performance liquid chromatography. Irradiation at 254 nm for 60 s in the presence of 1 mM [(14)C]halothane resulted in incorporation of approximately 0.5 mol of (14)C/mol of subunit, with photolabeling distributed within the nAChR extracellular and transmembrane domains, primarily at tyrosines. GammaTyr-111 in ACh binding site segment E was labeled, while alphaTyr-93 in segment A was not. Within the transmembrane domain, alphaTyr-213 within alphaM1 and deltaTyr-228 within deltaM1 were photolabeled, while no labeled amino acids were identified within the deltaM2 ion channel domain. Although the efficiency of photolabeling at the subunit level was unaffected by agonist, competitive antagonist, or isoflurane, state-dependent photolabeling was seen in a delta subunit fragment beginning at deltaPhe-206. Labeling of deltaTyr-212 in the extracellular domain was inhibited >90% by d-tubocurarine, whereas addition of either carbamylcholine or isoflurane had no effect. Within M1, the level of photolabeling of deltaTyr-228 with [(14)C]halothane was increased by carbamylcholine (90%) or d-tubocurarine (50%), but it was inhibited by isoflurane (40%). Within the structure of the nAChR transmembrane domain, deltaTyr-228 projects into an extracellular, water accessible pocket formed by amino acids from the deltaM1-deltaM3 alpha-helices. Halothane photolabeling of deltaTyr-228 provides initial evidence that halothane and isoflurane bind within this pocket with occupancy or access increased in the nAChR desensitized state compared to the closed channel state. Halothane binding at this site may contribute to the functional inhibition of nAChRs.  相似文献   

10.
Bupropion, a clinically used antidepressant and smoking-cessation drug, acts as a noncompetitive antagonist of nicotinic acetylcholine receptors (nAChRs). To identify its binding site(s) in nAChRs, we developed a photoreactive bupropion analogue, (±)-2-(N-tert-butylamino)-3'-[(125)I]-iodo-4'-azidopropiophenone (SADU-3-72). Based on inhibition of [(125)I]SADU-3-72 binding, SADU-3-72 binds with high affinity (IC(50) = 0.8 μM) to the Torpedo nAChR in the resting (closed channel) state and in the agonist-induced desensitized state, and bupropion binds to that site with 3-fold higher affinity in the desensitized (IC(50) = 1.2 μM) than in the resting state. Photolabeling of Torpedo nAChRs with [(125)I]SADU-3-72 followed by limited in-gel digestion of nAChR subunits with endoproteinase Glu-C established the presence of [(125)I]SADU-3-72 photoincorporation within nAChR subunit fragments containing M1-M2-M3 helices (αV8-20K, βV8-22/23K, and γV8-24K) or M1-M2 helices (δV8-14). Photolabeling within βV8-22/23K, γV8-24K, and δV8-14 was reduced in the desensitized state and inhibited by ion channel blockers selective for the resting (tetracaine) or desensitized (thienycyclohexylpiperidine (TCP)) state, and this pharmacologically specific photolabeling was localized to the M2-9 leucine ring (δLeu(265), βLeu(257)) within the ion channel. In contrast, photolabeling within the αV8-20K was enhanced in the desensitized state and not inhibited by TCP but was inhibited by bupropion. This agonist-enhanced photolabeling was localized to αTyr(213) in αM1. These results establish the presence of two distinct bupropion binding sites within the Torpedo nAChR transmembrane domain: a high affinity site at the middle (M2-9) of the ion channel and a second site near the extracellular end of αM1 within a previously described halothane (general anesthetic) binding pocket.  相似文献   

11.
Hamouda AK  Chiara DC  Blanton MP  Cohen JB 《Biochemistry》2008,47(48):12787-12794
The Torpedo nicotinic acetylcholine receptor (nAChR) is the only member of the Cys-loop superfamily of ligand-gated ion channels (LGICs) that is available in high abundance in a native membrane preparation. To study the structure of the other LGICs using biochemical and biophysical techniques, detergent solubilization, purification, and lipid reconstitution are usually required. To assess the effects of purification on receptor structure, we used the hydrophobic photoreactive probe 3-trifluoromethyl-3-(m-[(125)I]iodophenyl)diazirine ([(125)I]TID) to compare the state-dependent photolabeling of the Torpedo nAChR before and after purification and reincorporation into lipid. For the purified nAChR, the agonist-sensitive photolabeling within the M2 ion channel domain of positions M2-6, M2-9, and M2-13, the agonist-enhanced labeling of deltaThr274 (deltaM2-18) within the delta subunit helix bundle, and the labeling at the lipid-protein interface (alphaMu4) were the same as for the nAChR in native membranes. However, addition of agonist did not enhance [(125)I]TID photolabeling of deltaIle288 within the deltaM2-M3 loop. These results indicate that after purification and reconstitution of the Torpedo nAChR, the difference in structure between the resting and desensitized states within the M2 ion channel domain was preserved, but not the agonist-dependent change of structure of the deltaM2-M3 loop. To further characterize the pharmacology of [(125)I]TID binding sites in the nAChR in the desensitized state, we examined the effect of phencyclidine (PCP) on [(125)I]TID photolabeling. PCP inhibited [(125)I]TID labeling of amino acids at the cytoplasmic end of the ion channel (M2-2 and M2-6) while potentiating labeling at M2-9 and M2-13 and allosterically modulating the labeling of amino acids within the delta subunit helix bundle.  相似文献   

12.
The nicotinic acetylcholine receptor (nAChR) and the Na,K-ATPase functionally interact in skeletal muscle (Krivoi, I. I., Drabkina, T. M., Kravtsova, V. V., Vasiliev, A. N., Eaton, M. J., Skatchkov, S. N., and Mandel, F. (2006) Pflugers Arch. 452, 756–765; Krivoi, I., Vasiliev, A., Kravtsova, V., Dobretsov, M., and Mandel, F. (2003) Ann. N.Y. Acad. Sci. 986, 639–641). In this interaction, the specific binding of nanomolar concentrations of nicotinic agonists to the nAChR stimulates electrogenic transport by the Na,K-ATPase α2 isozyme, causing membrane hyperpolarization. This study examines the molecular nature and membrane localization of this interaction. Stimulation of Na,K-ATPase activity by the nAChR does not require ion flow through open nAChRs. It can be induced by nAChR desensitization alone, in the absence of nicotinic agonist, and saturates when the nAChR is fully desensitized. It is enhanced by noncompetitive blockers of the nAChR (proadifen, QX-222), which promote non-conducting or desensitized states; and retarded by tetracaine, which stabilizes the resting nAChR conformation. The interaction operates at the neuromuscular junction as well as on extrajunctional sarcolemma. The Na,K-ATPase α2 isozyme is enriched at the postsynaptic neuromuscular junction and co-localizes with nAChRs. The nAChR and Na,K-ATPase α subunits specifically coimmunoprecipitate with each other, phospholemman, and caveolin-3. In a purified membrane preparation from Torpedo californica enriched in nAChRs and the Na,K-ATPase, a ouabain-induced conformational change of the Na,K-ATPase enhances a conformational transition of the nAChR to a desensitized state. These results suggest a mechanism by which the nAChR in a desensitized state with high apparent affinity for agonist interacts with the Na,K-ATPase to stimulate active transport. The interaction utilizes a membrane-delimited complex involving protein-protein interactions, either directly or through additional protein partners. This interaction is expected to enhance neuromuscular transmission and muscle excitation.  相似文献   

13.
14.
The lipid state in acetylcholine receptor (AcChR)-rich membranes purified from electric organ of Torpedo marmorata was studied in the temperature interval from 0 degrees C to 35 degrees C using the (C-H) stretching and (C-C) skeletal optical vibrations. The Raman spectra of AcChR-rich membranes, recorded immediately after preparation of the samples, indicate that the lipids are in a predominant triclinic crystalline lattice and do not undergo a phase transition when the temperature increases up to 35 degrees C. However, the polar groups of the lipids appear subject to temperature-induced variations. After extraction of 43-kd and other non-receptor proteins, spectra indicate an order-disorder phase transition of lipids at approximately 21 degrees C. This transition appears less cooperative than the transition of the membrane lipid extract. The role of the proteins in preservation of the crystalline state of lipids in AcChR-rich membranes is discussed.  相似文献   

15.
A method for preparing thin, planar films of nicotinic acetylcholine receptor (nAChR) membranes that retain the ability to undergo the resting to desensitized state transition and that are suitable for spectroscopic studies has been developed. Native, alkaline-extracted nAChR membranes from Torpedo are dried under nitrogen on either a plastic microscope coverslip or a germanium internal reflection element (IRE) and then equilibrated with buffer. The drying procedure has no effect on the functional state of the nAChR as judged by a fluorescence assay using the probe ethidium bromide. The times required for an acetylcholine analogue (carbamylcholine), a local anesthetic (dibucaine), and a fluorescent probe (ethidium bromide) to penetrate films of varying degrees of thickness, interact with the receptor, and then to be washed from the films have been established. Under these conditions, the nAChR films can be repetitively cycled between the resting and desensitized states. Both fluorescence and infrared spectroscopy show that the films adhere strongly to either support even with buffer flowing continuously past the film surface. Fourier transform infrared difference spectra calculated from spectra recorded in the presence and absence of carbamylcholine show small, reproducible bands which reflect changes in nAChR structure upon desensitization.  相似文献   

16.
Stewart DS  Chiara DC  Cohen JB 《Biochemistry》2006,45(35):10641-10653
A molecule as simple in structure as tetramethylammonium gates the nicotinic acetylcholine receptor (nAChR) with high efficacy. To compare the structure of the nAChR transmitter binding site in the open channel state with that of the ACh binding protein, we determined the efficacy of nAChR gating by -S(CH(2))(n)N(CH(3))(3)(+) (n = 1-4) tethered to substituted cysteines at positions in the alpha subunits or gamma and delta subunits predicted to contribute to the ACh binding sites in mutant Torpedo nAChRs expressed in Xenopus oocytes. For tethered thiocholine [-S(CH(2))(2)N(CH(3))(3)(+)], we previously reported that within alpha195-201 gating was observed only at alphaY198C while at alphaY93C it acted as an antagonist. We now show that within alpha191-194, thiocholine activates when tethered at alphaCys192 or alphaCys193. Thiocholine also activates when tethered at alphaY190C or alphaW149C in nAChRs containing a beta subunit mutation (betaL257S) that destabilizes the closed channel, but not from gammaW55C/deltaW57C, where longer adducts can activate. When tethered at positions in binding site segment E, thiocholine activates only from gammaL119C/deltaL121C, where the shorter -S(CH(2))(1)N(CH3)(3)(+) acts as an antagonist. Longer adducts tethered at gammaL109C/deltaL111C or gammaL119C/deltaL121C also activate, but less efficiently. The length requirements for efficient gating by tethered agonists agree closely with predictions based upon the structure of the agonist site in a nAChR homology model derived from the ACh binding protein structure, which suggests that this structure is an excellent model of the nAChR agonist binding site in the open channel conformation. The inability of thiocholine to activate from alphaY93C, which is not predicted by the model, is discussed in terms of the structure of the nAChR in the closed state.  相似文献   

17.
Sticholysin I and II (St I and St II), two basic cytolysins purified from the Caribbean sea anemone Stichodactyla helianthus, efficiently permeabilize lipid vesicles by forming pores in their membranes. A general characteristic of these toxins is their preference for membranes containing sphingomyelin (SM). As a consequence, vesicles formed by equimolar mixtures of SM with phosphatidylcholine (PC) are very good targets for St I and II. To better characterize the lipid dependence of the cytolysin-membrane interaction, we have now evaluated the effect of including different lipids in the composition of the vesicles. We observed that at low doses of either St I or St II vesicles composed of SM and phosphatidic acid (PA) were permeabilized faster and to a higher extent than vesicles of PC and SM. As in the case of PC/SM mixtures, permeabilization was optimal when the molar ratio of PA/SM was ~1. The preference for membranes containing PA was confirmed by inhibition experiments in which the hemolytic activity of St I was diminished by pre-incubation with vesicles of different composition. The inclusion of even small proportions of PA into PC/SM LUVs led to a marked increase in calcein release caused by both St I and St II, reaching maximal effect at ~5 mol % of PA. Inclusion of other negatively charged lipids (phosphatidylserine (PS), phosphatidylglycerol (PG), phosphatidylinositol (PI), or cardiolipin (CL)), all at 5 mol %, also elicited an increase in calcein release, the potency being in the order CL approximately PA > PG approximately PI approximately PS. However, some boosting effect was also obtained, including the zwitterionic lipid phosphatidylethanolamine (PE) or even, albeit to a lesser extent, the positively charged lipid stearylamine (SA). This indicated that the effect was not mediated by electrostatic interactions between the cytolysin and the negative surface of the vesicles. In fact, increasing the ionic strength of the medium had only a small inhibitory effect on the interaction, but this was actually larger with uncharged vesicles than with negatively charged vesicles. A study of the fluidity of the different vesicles, probed by the environment-sensitive fluorescent dye diphenylhexatriene (DPH), showed that toxin activity was also not correlated to the average membrane fluidity. It is suggested that the insertion of the toxin channel could imply the formation in the bilayer of a nonlamellar structure, a toroidal lipid pore. In this case, the presence of lipids favoring a nonlamellar phase, in particular PA and CL, strong inducers of negative curvature in the bilayer, could help in the formation of the pore. This possibility is confirmed by the fact that the formation of toxin pores strongly promotes the rate of transbilayer movement of lipid molecules, which indicates local disruption of the lamellar structure.  相似文献   

18.
In this study, we investigated how the presence of anionic lipids influenced the stability and folding properties of the potassium channel KcsA. By using a combination of gel electrophoresis, tryptophan fluorescence and acrylamide quenching experiments, we found that the presence of the anionic lipid phosphatidylglycerol (PG) in a phosphatidylcholine (PC) bilayer slightly stabilized the tetramer and protected it from trifluoroethanol-induced dissociation. Surprisingly, the presence of phosphatidic acid (PA) had a much larger effect on the stability of KcsA and this lipid, in addition, significantly influenced the folding properties of the protein. The data indicate that PA creates some specificity over PG, and that it most likely stabilizes the tetramer via both electrostatic and hydrogen bond interactions.  相似文献   

19.
A M Haywood  B P Boyer 《Biochemistry》1984,23(18):4161-4166
How the lipid composition of liposomes determines their ability to fuse with Sendai virus membranes was tested. Liposomes were made of compositions designed to test postulated mechanisms of membrane fusion that require specific lipids. Fusion does not require the presence of lipids that can form micelles such as gangliosides or lipids that can undergo lamellar to hexagonal phase transitions such as phosphatidylethanolamine (PE), nor is a phosphatidylinositol (PI) to phosphatidic acid (PA) conversion required, since fusion occurs with liposomes containing phosphatidylcholine (PC) and any one of many different negatively charged lipids such as gangliosides, phosphatidylserine (PS), phosphatidylglycerol, dicetyl phosphate, PI, or PA. A negatively charged lipid is required since fusion does not occur with neutral liposomes containing PC and a neutral lipid such as globoside, sphingomyelin, or PE. Fusion of Sendai virus membranes with liposomes that contain PC and PS does not require Ca2+, so an anhydrous complex with Ca2+ or a Ca2+-induced lateral phase separation is not required although the possibility remains that viral binding causes a lateral phase separation. Sendai virus membranes can fuse with liposomes containing only PS, so a packing defect between domains of two different lipids is not required. The concentration of PS required for fusion to occur is approximately 10-fold higher than that required for ganglioside GD1a, which has been shown to act as a Sendai virus receptor. When cholesterol is added as a third lipid to liposomes containing PC and GD1a, the amount of fusion decreases if the GD1a concentration is low.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Using an acetylcholine-derivatized affinity column, we have purified human alpha4beta2 neuronal nicotinic acetylcholine receptors (nAChRs) from a stably transfected HEK-293 cell line. Both the quantity and the quality of the purified receptor are suitable for applying biochemical methods to directly study the structure of the alpha4beta2 nAChR. In this first study, the lipid-protein interface of purified and lipid-reconstituted alpha4beta2 nAChRs was directly examined using photoaffinity labeling with the hydrophobic probe 3-(trifluoromethyl)-3-(m-[125I]iodophenyl)diazirine ([125I]TID). [125I]TID photoincorporated into both alpha4 and beta2 subunits, and for each subunit the labeling was initially mapped to fragments containing the M4 and M1-M3 transmembrane segments. For both the alpha4 and beta2 subunits, approximately 60% of the total labeling was localized within fragments that contain the M4 segment, which suggests that the M4 segment has the greatest exposure to lipid. Within M4 segments, [125I]TID labeled homologous amino acids alpha4-Cys582/beta2-Cys445, which are also homologous to the [125I]TID-labeled residues alpha1-Cys418 and beta1-Cys447 in the lipid-exposed face of Torpedo nAChR alpha1M4 and beta1M4, respectively. Within the alpha4M1 segment, [125I]TID labeled residues Cys226 and Cys231, which correspond to the [125I]TID-labeled residues Cys222 and Phe227 at the lipid-exposed face of the Torpedo alpha1M1 segment. In beta2M1, [125I]TID labeled beta2-Cys220, which is homologous to alpha4-Cys226. We conclude from these studies that the alpha4beta2 nAChR can be purified from stably transfected HEK-293 cells in sufficient quantity and purity for structural studies and that the lipid-protein interfaces of the neuronal alpha4beta2 nAChR and the Torpedo nAChR display a high degree of structural homology.  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