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.
J W Nichols 《Biochemistry》1986,25(16):4596-4601
The rate of 1-palmitoyl-2-[12-[(7-nitro-2,1,3-benzoxadiazol-4-yl)amino] dodecanoyl] phosphatidylcholine (P-C12-NBD-PC) transfer between dioleoylphosphatidylcholine vesicles was measured by a technique based on resonance energy transfer between P-C12-NBD-PC and N-(lissamine rhodamine B sulfonyl)dioleoylphosphatidylethanolamine [Nichols, J. W., & Pagano, R. E. (1982) Biochemistry 21, 1720-1726]. Addition of bile salts at concentrations below their critical micelle concentrations increased the rate of spontaneous P-C12-NBD-PC transfer without disrupting the vesicles. The effectiveness in increasing the transfer rate was dependent on the structure of the bile salt. In general, conjugated bile salts were more effective than unconjugated, and mono- and dihydroxy bile salts were more effective than trihydroxy. The kinetics of intervesicular P-C12-NBD-PC transfer in the presence of cholate were found to be consistent with a mass action kinetic model based on the premise that bile salts bind to the vesicles, alter the dissociation and/or association rate constants for phospholipid monomer-vesicle interaction, and increase the rate of phospholipid transfer via the diffusion of soluble monomers through the aqueous phase. Temperature dependence studies indicated that cholate binding to vesicles is an entropy-driven process and that cholate binding lowers the free energy of activation for phospholipid monomer-vesicle dissociation by producing compensatory decreases in both the enthalpy and entropy of activation.  相似文献   

2.
The kinetics of spin-labeled phosphatidylcholine transfer between vesicles and HDL particles exhibited a two-phase process, as seen by ESR spectroscopy. The results were analyzed by considering several possible steps in the overall transfer, whose aspects were also studied: (i) micellar complex formation after HDL apolipoprotein-vesicle mixture, (ii) the rate of PC transfer from the micellar complex to HDL, (iii) the rate of the reverse reaction between overloaded HDL particles and other particles such as HDLs, LDLs, and lipid vesicles. The results agree most convincingly with a mechanism in which the diffusion of phospholipids into the HDL-endogenous lipids is the limiting step, occurring as a two-step process. In addition, we observed a negative charge effect on the lipid transfer rates and yields.  相似文献   

3.
J R Wu  B R Lentz 《Biochemistry》1991,30(27):6780-6787
Experiments were performed to assess three possible mechanisms of poly(ethylene glycol) (PEG) induced rapid lipid transfer between large unilamellar vesicles composed of dioleoylphosphatidylcholine: (1) transfer between aggregated vesicles, (2) transfer through an aqueous medium of lowered dielectric constant, and (3) transfer via a PEG carrier. The results showed that close contact between vesicles as a result of PEG dehydration was largely responsible for the rapid lipid transfer observed in the presence of PEG. The rate and extent of lipid transfer were also examined at 10 wt % PEG and analyzed in terms of a two-state model especially developed to account for the initial rate of lipid transfer as followed by the fluorescence lifetime of DPHpPC as a fluorescent lipid probe. Analysis revealed that two rate processes were involved in DPHpPC transfer between bilayers, both in the absence and presence of PEG. Since the maximum extent of transfer was 50%, transbilayer diffusion of DPHpPC seemed not to contribute to either process. The fast process in the presence of PEG was identified as due to rapid interbilayer monomer diffusion between closely apposed vesicles, and, in the absence of PEG, as due to monomer diffusion through the aqueous phase. The origin of the slow process, in both cases, remains obscure. The Arrhenius activation energies (and entropies) for the initial rates at temperatures from 10 to 48 degrees C were 15.3 +/- 0.3 kcal/mol (-26.3 +/- 0.2 eu) and 10.6 +/- 0.5 kcal/mol (-16.1 +/- 0.3 eu) in the absence and presence of PEG, respectively. The slow process was invariant with temperature.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The purpose of this report is to develop a correlation between the hydrophobicity of a phospholipid as measured by reversed-phase high-performance liquid chromatography and its rate of spontaneous transfer and to use this correlation to predict the rate of transfer of any homologous lipid from any lipoprotein. We have studied the mechanism of transfer of a series of fluorescent or radiolabeled phospholipids among natural and reassembled serum lipoproteins. Fluorescent phosphatidylcholines included those with 9-(1-pyrenyl)nonanoic acid in the sn-2 position and lauric, myristic, palmitic, stearic, oleic or linoleic acid at sn-1. The radioactive phosphatidylcholines contained [3H]oleic acid in the sn-2 position and lauric, myristic, or palmitic acid at sn-1. The kinetics of transfer of the pyrene-labeled lipid were followed by changes in the excimer fluorescence, and that of the radioactive lipids by separation of the donor (lipid-apolipoprotein recombinant) from the acceptor (single bilayer vesicles) on a column of Sephacryl S-200. The retention time of each lipid was measured by high-performance hydrophobic chromatography through a Waters radially compressed C18 column eluted with 75% isopropanol and 25% triethylammonium phosphate (0.15 M). A linear relationship was observed between the rate-constant of transfer and the retention time which suggest that the rate of desorption of phosphatidylcholines from lipoproteins and vesicles is controlled predominately by the hydrophobic effect. For a homologous series of lipids, the rate of transfer can be predicted from retention times obtained from hydrophobic chromatography. The kinetics of transfer of 1-lauroyl-2-[9-(1-pyrenyl)nonanoyl] phosphatidylcholine between isolated human serum lipoproteins exhibits a linear correlation between the transfer half-time and the size of the donor lipoproteins. As a consequence, transfer from very-low-density lipoprotein is 10-times slower than that observed from high-density lipoproteins. The observed correlations between phospholipid transfer rates and both the Stokes radius of the donor and the retention time of the phospholipid on a hydrophobic column permit one to calculate the rate of transfer of homologous molecules between lipid-protein complexes. The results predict that the spontaneous transfer of phospholipids between plasma lipoproteins would be too slow to be a physiologically important phenomena.  相似文献   

5.
Exchangeable phospho- and sphingolipid probes (phosphatidylcholine, -ethanolamine, -serine, and -glycerol, phosphatidic acid, sphingomyelin, cerebroside, and sulfatide) have been synthesized in which one acyl chain is substituted with a fluorescent bimanyl, 7-(dimethylamino)coumarin-3-yl, or diphenyl-hexatrienyl group. The distribution of these probes between two different populations of lipid vesicles can be readily monitored by fluorescence intensity measurements, as described by Nichols and Pagano [Nichols, J. W., & Pagano, R. E. (1982) Biochemistry 21, 1720-1726], when one of the vesicle populations contains a low mole fraction of a nonexchangeable quencher, (12-DABS)-18-PC. The probes examined in this study exchange between phospholipid vesicles on a time scale of minutes, with kinetics indicating that the transfer process takes place by diffusion of probe monomers through the aqueous phase. As expected, lipid probes with different charges differ markedly in their equilibrium distributions between neutral and charged lipid vesicles. However, probes with different polar headgroups differ only modestly in their relative affinities for vesicles composed of "hydrogen-bonding" lipids (PE and PS) vs "non-hydrogen-bonding" lipids (PC and PG or O-methyl-PA). Probes with different headgroups also show modest, albeit reproducible, differences in their relative affinities for cholesterol-containing vs cholesterol-free PC/PG vesicles. Our results suggest that lipids with different headgroup structures may mix more nearly ideally in liquid-crystalline lipid bilayers than would be predicted from previous analyses of the phase diagrams for binary lipid mixtures.  相似文献   

6.
Nonspecific lipid transfer proteins (nsLTPs) facilitate the transfer of phospholipids, glycolipids, fatty acids and steroids between membranes, with wide-ranging binding affinities. Three crystal structures of rice nsLTP1 from Oryza sativa, complexed with myristic (MYR), palmitic (PAL) or stearic acid (STE) were determined. The overall structures of the rice nsLTP1 complexes belong to the four-helix bundle folding with a long C-terminal loop. The nsLTP1-MYR and the nsLTP1-STE complexes bind a single fatty acid while the nsLTP1-PAL complex binds two molecules of fatty acids. The C-terminal loop region is elastic in order to accommodate a diverse range of lipid molecules. The lipid molecules interact with the nsLTP1-binding cavity mainly with hydrophobic interactions. Significant conformational changes were observed in the binding cavity and the C-terminal loop of the rice nsLTP1 upon lipid binding.  相似文献   

7.
The biosynthesis of nonspecific lipid transfer protein (nsLTP) was investigated. Total RNA of rat liver was translated in a rabbit reticulocyte lysate cell-free protein-synthesizing system with [35S]methionine as label. The immunoprecipitation of translation products with affinity-purified anti-nsLTP antibody yielded 14.5- and 60-kDa [35S]polypeptides. The molecular mass of the former polypeptide was approximately 1.5 kDa larger than that of the purified mature nsLTP (13 kDa). The site of synthesis of nsLTP was studied by in vitro translation of free and membrane-bound polyribosomal RNAs followed by immunoprecipitation. mRNA for both the 14.5- and 60-kDa polypeptides were found predominantly in the free polyribosomal fraction in both normal and clofibrate-treated rats. Clofibrate, a hypolipidemic drug that proliferates peroxisomes, did not increase the relative amount of nsLTP mRNA in rat liver. Pulse-chase experiments in rat hepatoma H-35 cells suggested that nsLTP was synthesized as a larger precursor of 14.5 kDa and converted to a mature form of 13 kDa. We have recently shown that nsLTP is highly concentrated in peroxisomes in rat hepatocytes [Tsuneoka et al. (1988) J. Biochem. 104, 560-564]. Taken together, these results suggest that nsLTP is synthesized as a larger precursor of 14.5 kDa on cytoplasmic free polyribosomes, then post-translationally transported to peroxisomes, where the precursor is presumably proteolytically processed to its mature form of 13 kDa. The relationship between the 13-kDa nsLTP and the 60-kDa polypeptide is also discussed.  相似文献   

8.
alpha-Sarcin is a fungal cytotoxic protein that inactivates the eukaryotic ribosomes. A kinetic study of the aggregation and lipid mixing promoted by this protein on phosphatidylglycerol (PG) and phosphatidylserine (PS) vesicles has been performed. Egg yolk PG, bovine brain PS, dimyristoyl-PG (DMPG) and dimyristoyl-PS (DMPS) vesicles have been considered. The initial rates of the vesicle aggregation induced by the protein have been measured by stopped-flow 90 degrees light scattering. The formation of a vesicle dimer as the initial step of this process was deduced from the second-order dependence of the initial rates on phospholipid concentration. The highest alpha-sarcin concentration studied did not inhibit the vesicle aggregation, indicating that many protein molecules are involved in the vesicle cross-linking. These are common characteristics of the initial steps of the aggregation produced by alpha-sarcin in the four types of phospholipid vesicles considered. However, the kinetics of the scattering values revealed that more complex changes occurred in the later steps of the aggregation process of egg PG and brain PS vesicles than in those of their synthetic counterparts. alpha-Sarcin produced lipid mixing in vesicles composed of DMPG or DMPS, which was measured by fluorescence resonance energy transfer assays. A delay in the onset of the process, dependent on the protein concentration, was observed. Measurement of the rates of lipid mixing revealed that the process is first order on phospholipid concentration. Egg PG and brain PS vesicles did not show lipid mixing, although they avidly aggregated. However, alpha-sarcin was able to promote lipid mixing in heterogeneous systems composed of egg PG+DMPG or brain PS+DMPS vesicles.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
V S Malinin  M E Haque  B R Lentz 《Biochemistry》2001,40(28):8292-8299
A number of fluorescent probes have been used to follow membrane fusion events, particularly intermixing of lipids. None of them is ideal. The most popular pair of probes is NBD-PE and Rh-PE, in which the fluorescent groups are attached to the lipid headgroups, making them sensitive to changes in the surrounding medium. Here we present a new assay for monitoring lipid transfer during membrane fusion using the acyl chain tagged fluorescent probes BODIPY500-PC and BODIPY530-PE. Like the NBD-PE/Rh-PE assay, this assay is based on fluorescence resonance energy transfer (FRET) between the donor, BODIPY500, and the acceptor, BODIPY530. The magnitude of FRET is sensitive to the probe surface concentration, allowing one to detect movement of probes from labeled to unlabeled vesicles during fusion. The high quantum yield of fluorescence, high efficiency of FRET (R(o) is estimated to be approximately 60 A), photostability, and localization in the central hydrophobic region of a bilayer all make this pair of probes quite promising for detecting fusion. We have compared this and two other lipid mixing assays for their abilities to detect the initial events of poly(ethylene glycol) (PEG)-mediated fusion of small unilamellar vesicles (SUVs). We found that the BODIPY500/530 assay showed lipid transfer rates consistent with those obtained using the DPHpPC self-quenching assay, while lipid mixing rates measured with the NBD-PE/Rh-PE RET assay were significantly slower. We speculate that the bulky labeled headgroups of NBD-PE and especially Rh-PE molecules hamper movement of probes through the stalk between fusing vesicles, and thus reduce the apparent rate of lipid mixing.  相似文献   

10.
The interfacial activation of porcine pancreatic phospholipase A(2) (PLA(2)) during the hydrolysis of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine liposomes at different temperatures has been monitored by fluorescence changes of the 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) lipid derivatives 1-palmitoyl-2-[6-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-glycero-3-phosphocholine (C(12)-NBD-PC) and 12-[(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)]dodecanoic acid (C(12)-NBD-FA) inserted in the substrate vesicles. These long-chain monitors, in contrast to the previously used C(6)-NBD-PC, detect latency times of PLA(2) action, similar to those measured by the classic titrimetric, pH-stat method. Interestingly, hydrolysis of the host vesicles results in a decrease in fluorescence not only of C(12)-NBD-PC, a substrate analog, but also of product derivative C(12)-NBD-FA. Ultrafiltration experiments show that C(12)-NBD-FA does not migrate to the aqueous phase upon hydrolysis of the host liposomes. Besides, in a simulated hydrolysis experiment in which increasing proportions of palmitic acid and 1-palmitoyl-sn-glycero-3-phosphocholine were cosonicated with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, C(12)-NBD-PC fluorescence was insensitive to products, whereas C(12)-NBD-FA did show a decreased emission intensity as in the actual hydrolysis experiments. The phenomenon is triggered above a critical concentration of products (10 mol%) suggesting that cosegregation of NBD-FA (either added as such or generated by hydrolysis of C(12)-NBD-PC) and products may be related to the decrease in fluorescence. Phase separation should create microdomains of increased C(12)-NBD-FA surface density and cause concentration quenching. In addition, and taking into account that the NBD group may be located near the interfacial region, it is possible that in segregating with products, the fluorescent moiety of C(12)-NBD-FA becomes exposed to microenvironments of higher surface polarity, which further decreases its quantum yield.  相似文献   

11.
Lin KF  Liu YN  Hsu ST  Samuel D  Cheng CS  Bonvin AM  Lyu PC 《Biochemistry》2005,44(15):5703-5712
Plant nonspecific lipid transfer proteins (nsLTPs) are thermal stable proteins that are capable of transferring lipid molecules between bilayers in vitro. This family of proteins, abundant in plants, is proposed to be involved in defense, pollination, and germination; the in vivo biological function remains, however, elusive. Here we report the purification and sequencing of an nsLTP1 from mung bean sprouts. We have also determined the solution structure of this nsLTP1, which represents the first 3D structure of the dicotyledonous nsLTP1 family. The global fold of mung bean nsLTP1 is similar to those of the monocotyledonous nsLTP1 structures and consists of four alpha-helices stabilized by four disulfide bonds. There are, however, some notable differences in the C-terminal tails and internal hydrophobic cavities. Circular dichroism and fluorescence spectroscopy were used to compare the thermodynamics and lipid transfer properties of mung bean nsLTP1 with those of rice nsLTP1. Docking of a lipid molecule into the solution structure of mung bean nsLTP1 reveals similar binding cavities and hydrophobic interactions as in rice nsLTP1, consistent with their comparable lipid transfer properties measured experimentally.  相似文献   

12.
In circulation the phospholipid transfer protein (PLTP) facilitates the transfer of phospholipid-rich surface components from postlipolytic chylomicrons and very low density lipoproteins (VLDL) to HDL and thereby regulates plasma HDL levels. To study the molecular mechanisms involved in PLTP-mediated lipid transfer, we studied the interfacial properties of PLTP using Langmuir phospholipid monolayers and asymmetrical flow field-flow fractionation (AsFlFFF) to follow the transfer of 14C-labeled phospholipids and [35S]PLTP between lipid vesicles and HDL particles. The AsFlFFF method was also used to determine the sizes of spherical and discoidal HDL particles and small unilamellar lipid vesicles. In Langmuir monolayer studies high-activity (HA) and low-activity (LA) forms of PLTP associated with fluid phosphatidylcholine monolayers spread at the air/buffer interphase. Both forms also mediated desorption of [14C]dipalmitoylphosphatidylcholine (DPPC) from the phospholipid monolayer into the buffer phase, even when it contained no physiological acceptor such as HDL. After the addition of HDL3 to the buffer, HA-PLTP caused enhanced lipid transfer to them. The particle diameter of HA-PLTP was approximately 6 nm and that of HDL3 approximately 8 nm as determined by AsFlFFF analysis. Using this method, it could be demonstrated that in the presence of HA-PLTP, but not LA-PLTP, [14C]DPPC was transferred from small unilamellar vesicles (SUV) to acceptor HDL3 molecules. Concomitantly, [35S]-HA-PLTP was transferred from the donor to acceptor, and this transfer was not observed for its low-activity counterpart. These observations suggest that HA-PLTP is capable of transferring lipids by a shuttle mechanism and that formation of a ternary complex between PLTP, acceptor, and donor particles is not necessary for phospholipid transfer.  相似文献   

13.
Cheng CS  Samuel D  Liu YJ  Shyu JC  Lai SM  Lin KF  Lyu PC 《Biochemistry》2004,43(43):13628-13636
Plant nonspecific lipid transfer proteins (nsLTPs) are small basic proteins that transport phospholipids between membranes. On the basis of molecular mass, nsLTPs are subdivided into nsLTP1 and nsLTP2. NsLTPs are all helical proteins stabilized by four conserved disulfide bonds. The existence of an internal hydrophobic cavity, running through the molecule, is a typical characteristic of nsLTPs that serves as the binding site for lipid-like substrates. NsLTPs are known to participate in plant defense, but the exact mechanism of their antimicrobial action against fungi or bacteria is still unclear. To trigger plant defense responses, a receptor at the plant surface needs to recognize the complex of a fungal protein (elicitin) and ergosterol. NsLTPs share high structural similarities with elicitin and need to be associated with a hydrophobic ligand to stimulate a defense response. In this study, binding of sterol molecules with rice nsLTPs is analyzed using various biophysical methods. NsLTP2 can accommodate a planar sterol molecule, but nsLTP1 binds only linear lipid molecules. Although the hydrophobic cavity of rice nsLTP2 is smaller than that of rice nsLTP1, it is flexible enough to accommodate the voluminous sterol molecule. The dissociation constant for the nsLTP2/cholesterol complex is approximately 71.21 microM as measured by H/D exchange and mass spectroscopic detection. Schematic models of the nsLTP complex structure give interesting clues about the reason for differential binding modes. Comparisons of NMR spectra of the sterol/rice nsLTP2 complex and free nsLTP2 revealed the residues involved in binding.  相似文献   

14.
The non‐specific lipid transfer proteins (nsLTPs) are multifunctional seed proteins engaged in several different physiological processes. The nsLTPs are stabilized by four disulfide bonds and exhibit a characteristic hydrophobic cavity, which is the primary lipid binding site. While these proteins are known to transfer lipids between membranes, the mechanism of lipid transfer has remained elusive. Four crystal structures of nsLTP from Solanum melongena, one in the apo‐state and three myristic acid bound states were determined. Among the three lipid bound states, two lipid molecules were bound on the nsLTP surface at different positions and one was inside the cavity. The lipid‐dependent conformational changes leading to opening of the cavity were revealed based on structural and spectroscopic data. The surface‐bound lipid represented a transient intermediate state and the lipid ultimately moved inside the cavity through the cavity gate as revealed by molecular dynamics simulations. Two critical residues in the loop regions played possible ‘gating’ role in the opening and closing of the cavity. Antifungal activity and membrane permeabilization effect of nsLTP against Fusarium oxysporum suggested that it could possibly involve in bleaching out the lipids. Collectively, these studies support a model of lipid transfer mechanism by nsLTP via intermediate states.  相似文献   

15.
The genes encoding non-specific lipid transfer proteins (nsLTPs), members of a small multigene family, show a complex pattern of expressional regulation, suggesting that some diversification may have resulted from changes in their expression after duplication. In this study, the evolution of nsLTP genes within the Poaceae family was characterized via a survey of the pseudogenes and unigenes encoding the nsLTP in rice pseudomolecules and the NCBI unigene database. nsLTP-rich regions were detected in the distal portions of rice chromosomes 11 and 12; these may have resulted from the most recent large segmental duplication in the rice genome. Two independent tandem duplications were shown to occur within the nsLTP-rich regions of rice. The genomic distribution of the nsLTP genes in the rice genome differs from that in wheat. This may be attributed to gene migration, chromosomal rearrangement, and/or differential gene loss. The genomic distribution pattern of nsLTP genes in the Poaceae family points to the existence of some differences among cereal nsLTP genes, all of which diverged from an ancient gene. The unigenes encoding nsLTPs in each cereal species are clustered into five groups. The somewhat different distribution of nsLTP-encoding EST clones between the groups across cereal species imply that independent duplication(s) followed by subfunctionalization (and/or neofunctionalization) of the nsLTP gene family in each species occurred during speciation.  相似文献   

16.
C Pryor  M Bridge  L M Loew 《Biochemistry》1985,24(9):2203-2209
A new fluorescent lipid analogue, bimanephosphatidylcholine, has been synthesized for use in lipid bilayers. This probe is well suited as an energy-transfer donor with N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine as the acceptor. Dimyristoylphosphatidylethanolamine vesicles are prepared by sonication at pH 9 and characterized by electron microscopy and other methods. Resonance energy transfer between separately labeled donor and acceptor vesicles is monitored during HCl-induced aggregation to determine the kinetics of lipid randomization. Light scattering is also monitored to measure the kinetics of aggregation. The light scattering shows a marked reversal with NaOH while the energy transfer does not, indicating lipid exchange during a reversibly aggregated state; the extent of energy transfer suggests that only lipids in the outer monolayers exchange. The gel to liquid-crystalline phase transition temperature in HCl-treated vesicles is found to be 47 degrees C with diphenylhexatriene. The initial sonicated dispersion does not show a sharp phase transition. In vesicles labeled with both donor and acceptor probes, a small, irreversible increase in energy transfer is obtained upon lowering and then restoring the pH. These results suggest a metastable phase in the sonicated vesicles containing a randomized distribution of lipid and probes within the bilayers; the thermodynamically favored phase, whose formation is triggered by the pH shock, contains domains within which the probe lipids are more highly concentrated.  相似文献   

17.
We have examined the internalization and degradation of a fluorescent analog of phosphatidylcholine after its insertion into the plasma membrane of cultured Chinese hamster fibroblasts. 1-acyl-2-(N-4- nitrobenzo-2-oxa-1,3-diazole)-aminocaproyl phosphatidylcholine (C6-NBD- PC) was incorporated into the cell surface by liposome-cell lipid transfer at 2 degrees C. The fluorescent lipid remained localized at the plasma membrane as long as the cells were kept at 2 degrees C; however, when the cells were warmed to 37 degrees C, internalization of some of the fluorescent lipid occurred. Most of the internalized C6-NBD- PC accumulated in the Golgi apparatus although a small amount was found randomly distributed throughout the cytoplasm in punctate fluorescent structures. Internalization of the fluorescent lipid at 37 degrees C was blocked by the presence of inhibitors of endocytosis. Incubation of cells containing C6-NBD-PC at 37 degrees C resulted in a rapid degradation of the fluorescent lipid. This degradation occurred predominantly at the plasma membrane. The degradation of C6-NBD-PC resulted in the release of NBD-fatty acid into the medium. We have compared the internalization of the fluorescent lipid with that of a fluorescent protein bound to the cell surface. Both fluorescent lipid and protein remained at the plasma membrane at 2 degrees C and neither were internalized at 37 degrees C in the presence of inhibitors of endocytosis. However, when incubated at 37 degrees C under conditions that permit endocytosis, the two fluorescent species appeared at different intracellular sites. Our data suggest that there is no transmembrane movement of C6-NBD-PC and that the fluorescent probe reflects the internalization of the outer leaflet of the plasma membrane lipid bilayer. The results are consistent with the Golgi apparatus as being the primary delivery site of phospholipid by bulk membrane movement from the plasma membrane.  相似文献   

18.
The kinetics of poly(ethylene glycol) (PEG)-induced fusion between intact human erythrocytes was continuously monitored by a fluorescence lipid mixing method, utilizing the dequenching of the fluorescence probe, 1-oleoyl-2-[12-[(7-nitro-2,1,3-benzoxadiazol-4-yl)amino]dodecanoyl ] phosphatidylcholine (C12-NBD-PC). The steady-state fluorescence intensity was detected from the surface of cells in a monolayer on an alcian blue-coated glass coverslip. The relief of fluorescence self-quenching after fusion between C12-NBD-PC labeled and unlabeled intact erythrocytes was measured. The extent of fluorescence dequenching was normalized based on the measured concentration of probes in membranes, the projected partial dequenching due both to dilution by intercellular fusion, and the dilution between the inner and outer leaflets of membranes (flip-flop). There was no significant increase in fluorescence intensity during PEG treatment of 5 min, at 4 degrees C. Intensity increased immediately after the dilution of PEG, and reached saturation in 30 min. The efficiency of fusion increased with the increasing of PEG concentrations. Only 4% enhancement of saturated relative fluorescence intensity was detected in 25 wt% PEG-induced cell fusion; 23% enhancement in 30 wt%; and 66% enhancement in 35 wt%. The transfer of fluorescent probes between membrane bilayer leaflets (flip-flop) was also monitored during the fusion process. Flip-flop was monitored in confluent monolayers as well as in isolated cells. There was no significant spontaneous flip-flop within 30 min of dilution. The relative fluorescence intensity enhancement contributed by the dilution of probes between fused labeled and unlabeled cells (at a 1:1 ratio) was found to account for only 39% of the observed final dequenching, whereas the contribution by flip-flop associated with cell fusion was found to account for 9%, and flip-flop without fusion contributed approximately 18%. A portion of the flip-flop is a consequence of hemolysis. Therefore, fluorescence dequenching measurements of fusion of whole cells must be interpreted with caution.  相似文献   

19.
Nonspecific lipid transfer protein (sterol carrier protein2) has previously been proposed to function as (i) a catalyst for intracellular movement of newly synthesized phospholipid, (ii) a cofactor in the biosynthesis and metabolism of cholesterol, and (iii) a cofactor in the feedback inhibition of cholesterol synthesis. Each of these functions is based upon the premise that nonspecific lipid transfer protein (nsLTP) is cytosolic. However, evidence presented in this report suggests that, at least in the case of cultured hepatoma cells, nsLTP is secreted. This conclusion is supported by three observations. First, after culture of hepatoma cells for 10 h, 88% of the nsLTP (as judged by its phosphatidylethanolamine transfer activity) appears in the medium, whereas the cytosolic level of transfer activity remains unchanged. Furthermore, this is accompanied by the appearance in the medium of a polypeptide of Mr 12,200-12,500, which corresponds to the known molecular weight of nsLTP. Finally, it was observed that the appearance of both the activity and the polypeptide in the medium are inhibited by monensin, an inhibitor of secretion. Thus their appearance seems to represent secretion and not simply leakage from the cells. Further evidence that nsLTP does not play an important role in the cytosolic transport of phospholipid and sterol is provided by our observation that hepatoma cells containing a level of nsLTP only 10-15% of that found in liver nevertheless possess near-normal membrane phospholipid compositions and retain the ability to feedback-inhibit cholesterol biosynthesis.  相似文献   

20.
Rao CS  Lin X  Pike HM  Molotkovsky JG  Brown RE 《Biochemistry》2004,43(43):13805-13815
Glycolipid transfer protein (GLTP) catalyzes the intermembrane transfer of lipids that have sugars beta-linked to either diacylglycerol or ceramide backbones, including simple glycosphingolipids (GSLs) and gangliosides. The present study provides a quantitative understanding of GLTP action involving bilayer vesicles that have high and low curvature stress, i.e., small and large unilamellar vesicles (SUVs and LUVs). When the GSL intervesicular transfer was monitored in real time using an established fluorescence resonance energy approach, the initial GSL transfer rates (v(0)) and net transfer equilibrium (K(eq)) were determined for GLTP-mediated transfer from SUVs and LUVs over the temperature range of 30-44 degrees C. v(0) exhibited a linear dependence with respect to varying GLTP concentrations (0-143 nM range) in SUVs and LUVs, suggesting a first order dependence on the GLTP bulk concentration. Thermodynamic parameters associated with the GLTP-GSL transition-state complex and GSL net transfer were determined from linear Arrhenius and van't Hoff plots, respectively. Although initial transfer rates were lower for LUVs than for SUVs, the activation energy barriers were higher for LUVs, while the Gibbs's free energy of the transition states were similar. The formation of a transition-state complex was predominantly enthalpy driven, whereas the net transfer of GSLs was mainly entropy driven. The rate-limiting step for GLTP action appeared to be the surface processes leading to the GLTP-GSL complex formation and release associated with a shuttle/carrier mode of action. Because surface processes leading to the GLTP-GSL complex formation were limiting for GLTP action with SUVs and LUVs, it was concluded that GLTP is likely to be a valuable tool to probe and manipulate GSL environments in membranes.  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