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 interfacial elastic packing interactions of different galactosylceramides (GalCers), sphingomyelins (SMs), and phosphatidylcholines (PC) were compared by determining their elastic area compressibility moduli (Cs-1) as a function of lateral packing pressure (pi) in a Langmuir-type film balance. To assess the relative contributions of the lipid headgroups as well as those of the ceramide and diacylglycerol hydrocarbon regions, we synthesized various GalCer and SM species with identical, homogeneous acyl residues and compared their behavior to that of PCs possessing similar hydrocarbon structures. For PCs, this meant that the sn-1 acyl chain was long and saturated (e.g., palmitate) and the sn-2 chain composition was varied to match that of GalCer or SM. When at equivalent pi and in either the chain-disordered (liquid-expanded) or chain-ordered (liquid-condensed) state, GalCer films were less elastic than either SM or PC films. When lipid headgroups were identical (SM and PC), Cs-1 values (at equivalent pi) for chain-disordered SMs, but not chain-ordered SMs, were 25-30% higher than those of PCs. Typical values for fluid phase (liquid-expanded) GalCer at 30 mN/m and 24 degrees C were 158 (+/- 7) mN/m, whereas those of SM were 135 (+/- 7) mN/m and those of PC were 123 (+/- 2) mN/m. Pressure-induced transitions to chain-ordered states (liquid-condensed) resulted in significant increases (two- to fourfold) in the "in-plane" compressibility for all three lipid types. Typical Cs-1 values for chain-ordered GalCers at 30 mN/m and 24 degrees C were between 610 and 650 mN/m, whereas those of SM and of PC were very similar and were between 265 and 300 mN/m. Under fluid phase conditions, the pi-Cs-1 behavior for each lipid type was insensitive to whether the acyl chain was saturated or monounsaturated. Measurement of the Cs-1 values also provided an effective way to evaluate the two-dimensional phase transition region of SMs, GalCers, and PCs. Modest heterogeneity in the acyl composition led to transitional broadening. Our findings provide useful information regarding the in-plane elasticity of lipids that are difficult to investigate by alternative methods, i.e., micropipette aspiration technique. The results also provide insight into the stability of sphingolipid-enriched, membrane microdomains that are thought to play a role in the sorting and trafficking of proteins containing glycosylphosphatidylinositol anchors with cells.  相似文献   

2.
We have examined the association of 5-androsten-3beta-ol (androsterol) with saturated phosphatidylcholines (PCs), having symmetric acyl chains from 10 to 16 carbons in length, in both mono- and bilayer membranes. The emphasis of the study was to measure how hydrophobic mismatch (i.e. the difference in hydrophobic length of the interacting molecules) affected androsterol/PC interactions in model membranes. With monolayer membranes (33 mol% sterol, 20 mN/m, 25 degreesC), androsterol was found to be macroscopically miscible with all the tested PCs. Androsterol was observed to condense the lateral packing of di14 and di15 PCs (by 6 and 4.5 A2 per molecule, respectively), but failed to condense shorter (di10, di11, di12 and di13 PCs) or the longer chain di16PC. The rate of androsterol desorption from mixed monolayers to beta-cyclodextrin acceptors in the subphase was a clear function of the host PC acyl chain length. The slowest rate of androsterol desorption (i.e. best androsterol/PC interaction) was seen from a di14PC monolayer, whereas the desorption rate increased when the host PC had shorter or longer chains. When the cholesterol oxidase susceptibility of androsterol was determined in small unilamellar vesicles (SUV) containing PCs of different chain lengths (33 mol% androsterol), the slowest rate of oxidation was seen in di14PC vesicles, whereas higher rates were measured for shorter or longer chain PC vesicles, again suggesting that androsterol interacted more favorably with di14PC than with the other PCs. In conclusion, the hydrophobic mismatch between androsterol and different PCs appeared to greatly affect the intermolecular interactions, as determined from the condensation effect, from sterol desorption rates, and the oxidation susceptibility of androsterol. Although androsterol is not a physiological membrane component, the present model system clearly shows that hydrophobic mismatch has a great influence on how sterols and phosphatidylcholines interact in membranes.  相似文献   

3.
Phosphatidylcholines (PCs) with stearoyl (18:0) sn-1 chains and variable-length, saturated sn-2 acyl chains were synthesized and investigated using a Langmuir-type film balance. Surface pressure was monitored as a function of lipid molecular area at various constant temperatures between 10 degrees C and 30 degrees C. Over this temperature range, 18:0-10:0 PC displayed only liquid-expanded behavior. In contrast, di-14:0 PC displayed liquid-expanded behavior at 24 degrees C and 30 degrees C, but two-dimensional phase transitions were evident at 20 degrees C, 15 degrees C, and 10 degrees C. The average molecular area of 18:0-10:0 PC was larger than that of liquid-expanded di-14:0 PC at equivalent surface pressures, and the shapes of their liquid expanded isotherms were somewhat dissimilar. Analysis of the elastic moduli of area compressibility (Cs(-1)) as a function of molecular area revealed shallower slopes in the semilog plots of 18:0-10:0 PC compared to di-14:0 PC. At membrane-like surface pressures (e.g., 30 mN/m), 18:0-10:0 PC was 20-25% more elastic (in an in-plane sense) than di-14:0 PC. Other PCs with varying degrees of chain-length asymmetry (18:0-8:0 PC, 18:0-12:0 PC, 18:0-14:0 PC, 18:0-16:0 PC) were also investigated to determine whether the higher in-plane elasticity of fluid-phase 18:0-10:0 PC is a common feature of PCs with asymmetrical chain lengths. Two-dimensional phase transitions in 18:0-14:0 PC and 18:0-16:0 PC prevented meaningful comparison with other fluid-phase PCs at 30 mN/m. However, the Cs(-1) values for fluid-phase 18:0-8:0 PC and 18:0-12:0 PC were similar to that of 18:0-10:0 PC (85-90 mN/m). These values showed chain-length asymmetrical PCs to have 20-25% greater in-plane elasticity than fluid-phase PCs with mono- or diunsaturated acyl chains.  相似文献   

4.
Although dietary trans unsaturated fatty acids (TUFA) are known to decrease plasma HDL, the underlying mechanisms for this effect are unclear. We tested the hypothesis that the decreased HDL is due to an inhibition of lecithin:cholesterol acyltransferase (LCAT), the enzyme essential for the formation of HDL, by determining the activity of purified LCAT in the presence of synthetic phosphatidylcholine (PC) substrates containing TUFA. Both human and rat LCATs exhibited significantly lower activity (-37% to -50%) with PCs containing 18:1t or 18:2t, when compared with the PCs containing corresponding cis isomers. TUFA-containing PCs also inhibited the enzyme activity competitively, when added to egg PC substrate. The inhibition of LCAT activity was not due to changes in the fluidity of the substrate particle. However, the inhibition depended on the position occupied by TUFA in the PC, as well as on the paired fatty acid. Thus, for human LCAT, 18:1t was more inhibitory when present at sn-2 position of PC, than at sn-1, when paired with 16:0. In contrast, when paired with 20:4, 18:1t was more inhibitory at sn-1 position of PC. Both human and rat LCATs, which are normally specific for the sn-2 acyl group of PC, exhibited an alteration in their positional specificity when 16:0-18:1t PC or 16:1t-20:4 PC was used as substrate, deriving 26-86% of the total acyl groups for cholesterol esterification from the sn-1 position. These results show that the trans fatty acids decrease high density lipoprotein through their inhibition of lecithin: cholesterol acyltransferase (LCAT) activity, and also alter LCAT's positional specificity, inducing the formation of more saturated cholesteryl esters, which are more atherogenic.  相似文献   

5.
A series of diaryl-substituted heterocyclic ureas was prepared, and their ability to inhibit acyl-CoA: cholesterol O-acyltransferase (ACAT) in vitro and to lower plasma total cholesterol in cholesterol-fed animal models in vivo was examined. N-(2,6-Diisopropylphenyl)-N'-tetrazole or isoxazole-substituted heterocyclic ureas proved optimal. A carbon chain of 11-14 carbons substituted 1,3 with respect to the amine provided the optimal side chain. Substitution of the alkyl chain generally lowered activity. Tetrazole urea 2i dosed at 3 mg/kg lowered plasma total cholesterol (TC) 67% in an acute, cholesterol-fed (C-fed) rat model of hypercholesterolemia and 47% in C-fed dogs. Tetrazole 2i, dosed at 10 mg/kg, also lowered TC 52% and raised HDL cholesterol 113% in rats with pre-established hypercholesterolemia.  相似文献   

6.
We have determined calorimetrically the phase transition temperature (Tm) values of five sn-1 saturated/sn-2 unsaturated phosphatidylethanolamines (PE) in which the sn-1 acyl chain has 20 carbons and the sn-2 acyl chain has 18 carbons with different number and position of the cis double bond. When these Tm values are combined with the five published Tm values of related unsaturated PE, a unifying Tm diagram is generated for the first time. Moreover, as the molecular mechanics simulated structures of these lipids are taken into consideration, this unifying Tm diagram provides insight into how variations in the number and position of the cis double bond in the lipid's sn-2 acyl chain can influence the phase transition behavior of the lipid bilayer.  相似文献   

7.
The effects of polyunsaturated fatty acids (PUFA) on the structure of recombinant high density lipoprotein (rHDL) was investigated using homogeneous particles containing phosphatidylcholine (PC), [3H]cholesterol, and apolipoprotein A-I (apoA-I). The PC component of the rHDL contained sn -1 16:0 and sn -2 18:1 (POPC), 18:2 (PLPC), 20:4 (PAPC), 20:5 n-3 (PEPC), or 22:6 n-3 (PDPC). The concentration of guanidine HCl (D1/2) required to denature one-half of the apoA-I on rHDL containing long chain PUFA was reduced (1.57-1.70 m) compared to those containing POPC (2.83 m). Intrinsic apoA-I tryptophan fluorescence emission intensity and lifetimes were decreased for rHDL containing long chain PUFA compared to POPC and PLPC rHDL. Monoclonal antibody binding studies demonstrated that apoA-I had decreased immunoreactivity with monoclonal antibodies spanning amino acid residues 115-147 in rHDL containing long chain PUFA. PC lipid fluidity, measured as diphenylhexatriene (DPH) fluorescence polarization, was increased in PUFA rHDL compared to POPC rHDL. There also was a strong correlation between the number of sn -2 double bonds in rHDL and DPH fluorescence lifetime (r 2 = 0. 89). LCAT reactivity of the homogeneous size rHDL was ordered POPC = PLPC>PAPC> PEPC>PDPC. We conclude that rHDL with long chain PUFA in the sn -2 position of PC contain apoA-I that is less stable and in a different conformation than that in POPC rHDL and have a fatty acyl region that is more fluid and hydrated. The weaker interaction of apoA-I with PC containing PUFA may lead to hypercatabolism of apoA-I in plasma explaining, in part, the decreased plasma HDL and apoA-I concentrations seen with PUFA diets.  相似文献   

8.
African green monkeys fed fat-specific diets served as a model to investigate the effect of phospholipid acyl chain modification on high density lipoprotein (HDL)-mediated cellular cholesterol efflux. Diets enriched in saturated, monounsaturated, n-6 polyunsaturated, or n-3 polyunsaturated fats were provided during both low cholesterol and cholesterol-enriched stages; sera and HDL3 samples were obtained at specific points during the treatment period. Analysis of the HDL phospholipid composition revealed significant acyl chain modification, consistent with the respective fat-specific diet. Cholesterol efflux from mouse L-cell fibroblasts to HDL3 isolated from the specific diet groups was measured and revealed no differences in the abilities of the particles to accept cellular cholesterol; determination of the bidirectional flux of cholesterol between the cells and HDL3 species further demonstrated no effect of phospholipid acyl chain modification on this process. The effects of dietary modification of phospholipid acyl chains on cellular cholesterol efflux were directly examined by isolating the HDL phospholipid and combining it with human apolipoprotein A-I to form well-defined reconstituted HDL particles. These complexes did not display any differences with respect to their ability to stimulate cellular cholesterol efflux. Incubations with 5% sera further confirmed that the fat-specific diets do not influence cholesterol efflux. These results suggest that the established influences of specific dietary fats on the progression of atherosclerosis are due to effects on cholesterol metabolism other than the efflux of cellular cholesterol in the first step of reverse cholesterol transport.  相似文献   

9.
Several novel N-(4,5-diphenylthiazol-2-yl)-N'-aryl or alkyl (thio)ureas and N-(4,5-diphenylthiazol-2-yl)alkanamides were prepared as potential acyl-CoA: cholesterol O-acyltransferase (ACAT) inhibitors. Synthesis was accomplished by reaction of 2-amino-4,5-diphenylthiazole with the suitable isocyanate, isothiocyanate or acyl chloride. Some analogues without the 5-phenyl substituent or both the phenyl groups in 4 and 5 position of the thiazole ring were also prepared. Moreover, some bioisosters of the title compounds in which the thiazole ring was replaced by an imidazole were synthesized starting from the 2-amino-4,5-diphenyl-1H-imidazole. The ability of synthesized compounds to inhibit ACAT was evaluated in vitro by measuring the formation of cholesteryl[14C]oleate from cholesterol and [1-14C]oleoyl-CoA in rat liver microsomes. Among the tested compounds, only some thiazole ureas were able to inhibit ACAT in a reasonable degree. N-(4,5-diphenylthiazol-2-yl)- N'-[2,6-bis(2-methylethyl)phenyl] urea (11) and N-(4,5-diphenylthiazol-2-yl)-N'-n-butyl urea (16) were the most active compounds in the series showing IC50 values in the low micromolar range.  相似文献   

10.
We have shown previously that lateral compression of pulmonary surfactant monolayers initially induces separation of two phases but that these remix when the films become more dense (1). In the studies reported here, we used fluorescence microscopy to examine the role of the different surfactant constituents in the remixing of the separated phases. Subfractions containing only the purified phospholipids (PPL), the surfactant proteins and phospholipids (SP&PL), and the neutral and phospholipids (N&PL) were obtained by chromatographic separation of the components in extracted calf surfactant (calf lung surfactant extract, CLSE). Compression of the different monolayers produced nonfluorescent domains that emerged for temperatures between 20 and 41 degreesC at similar surface pressures 6-8 mN/m higher than values observed for dipalmitoyl phosphatidylcholine (DPPC), the most prevalent component of pulmonary surfactant. Comparison of the different preparations showed that the neutral lipid increased the total nonfluorescent area at surface pressures up to 25 mN/m but dispersed that total area among a larger number of smaller domains. The surfactant proteins also produced smaller domains, but they had the opposite effect of decreasing the total nonfluorescent area. Only the neutral lipids caused remixing. In images from static monolayers, the domains for N&PL dropped from a maximum of 26 +/- 3% of the interface at 25 mN/m to 4 +/- 2% at 30 mN/m, similar to the previously reported behavior for CLSE. During continuous compression through a narrow range of pressure and molecular area, in N&PL, CLSE, and mixtures of PPL with 10% cholesterol, domains became highly distorted immediately prior to remixing. The characteristic transition in shape and abrupt termination of phase coexistence indicate that the remixing caused by the neutral lipids occurs at or close to a critical point.  相似文献   

11.
The structure and thermotropic properties of N-palmitoyl sphingomyelin (C16:0-SM) and its interaction with cholesterol and dipalmitoylphosphatidylcholine (DPPC) have been studied by differential scanning calorimetry (DSC) and X-ray diffraction methods. DSC of hydrated multi-bilayers of C16:0-SM shows reversible chain-melting transitions. On heating, anhydrous C16:0-SM exhibits an endothermic transition at 75 degrees C (delta H = 4.0 kcal/mol). Increasing hydration progressively lowers the transition temperature (TM) and increases the transition enthalpy (delta H), until limiting values (TM = 41 degrees C, delta H = 7.5 kcal/mol) are observed for hydration values > 25 wt % H2O. X-ray diffraction at temperatures below (29 degrees C) TM show a bilayer gel structure (d = 73.5 A, sharp 4.2 A reflection) for C16:0-SM at full hydration; above TM, at 55 degrees C, a bilayer liquid-crystal phase is present (d = 66.6 A, diffuse 4.6 A reflection). Addition of cholesterol to C16:0-SM bilayers results in a progressive decrease in the enthalpy of the transition at 41 degrees C, and no cooperative transition is detected at > 50 mol % cholesterol. X-ray diffraction shows no difference in the bilayer periodicity, position/width of the wide-angle reflections, or electron density profiles at 29 and 55 degrees C when 50 mol % cholesterol is present. Thus, cholesterol inserts into C16:0-SM bilayers progressively removing the chain-melting transition and changing the structural characteristics of the bilayer. DSC and X-ray diffraction data show that DPPC is completely miscible with C16:0-SM bilayers in both the gel and liquid-crystalline phases; however, 30 mol % C16:0-SM removes the pre-transition exhibited by DPPC.  相似文献   

12.
The behaviors of two chemically well-defined sphingolipids, N-palmitoyl-sphingomyelin (C16:0-SM) and the corresponding ceramide (C16:0-Cer), in a 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) matrix were compared. Minor attenuation of lateral diffusion upon increasing the mole fraction of C16:0-SM (XSM, up to 0.25) was indicated by the slight decrement in the excimer/monomer intensity ratio (Ie/Im) for a trace amount (mole fraction X = 0.01) of a pyrene-labeled ceramide analogue (N-[(pyren)-1-yl]decanoyl-sphingosine, PDCer) in keeping with the miscibility of C16:0-SM in POPC. Increasing membrane order was revealed by the augmented polarization P for diphenylhexatriene (DPH). In contrast, when C16:0-Cer was substituted for C16:0-SM an approximately 1.6-fold increase in Ie/Im for PDCer was evident upon increasing Xcer, with parallel increment in DPH polarization. In agreement with our recent data on natural ceramides in dimyristoylphosphatidylcholine (DMPC) bilayers [Holopainen et al. (1997) Chem. Phys. Lipids 88, 1-13], we conclude that C16:0-Cer becomes enriched into microdomains in the fluid POPC membrane. Interestingly, enhanced formation of microdomains by ceramide was observed when the total sphingolipid content in tertiary alloys with POPC was maintained constant (Xcer + XSM = 0.25) and the SM/Cer stoichiometry was varied. Finally, when ceramide was generated enzymatically in POPC/C16:0-SM (3:1, molar fraction) LUVs by sphingomyelinase (SMase, Bacillus cereus), maximally approximately 85% of hydrolysis of sphingomyelin was measured within <3 min at 30 degreesC. The formation of ceramide was accompanied by a closely parallel increase in DPH polarization. There was also an increase in Ie/Im for PDCer; however, these changes in Ie/Im were significantly slower, requiring approximately 105 min to reach a steady state. These data show that the rapid enzymatic formation of ceramide under these conditions is followed by much slower reorganization process, resulting in the formation of microdomains enriched in this lipid.  相似文献   

13.
We have investigated the mechanisms of interaction of the recombinant N-terminal portion of bactericidal/permeability-increasing protein, rBPI21, with lipopolysaccharide (LPS) isolated from enterobacterial deep rough mutant strains. Experimentally, the ability of rBPI21 to form monolayers at the air/water interface and its action on lipid monolayers were analyzed. We have further studied the interaction of rBPI21 with aggregates from phospholipids and Re mutant LPS by infrared and resonance energy transfer spectroscopy and laser Doppler velocimetry. From monolayer experiments, the molecular area of a single rBPI21 molecule was estimated to be about 12 nm2. At lateral pressures of 相似文献   

14.
We investigated in vivo expression of myosin heavy chain (MHC) isoforms, 17 kDa myosin light chain (MLC17), and phosphorylation of the 20 kDa MLC (MLC20) as well as mechanical performance of chemically skinned fibers of normal and hypertrophied smooth muscle (SM) of human myometrium. According to their immunological reactivity, we identified three MHC isoenzymes in the human myometrium: two SM-MHC (SM1 with 204 kDa and SM2 with 200 kDa), and one non-muscle specific MHC (NM with 196 kDa). No cross-reactivity was detected with an antibody raised against a peptide corresponding to a seven amino acid insert at the 25K/50K junction of the myosin head (a-25K/50K) in both normal and hypertrophied myometrium. In contrast, SM-MHC of human myomatous tissue strongly reacted with a-25K/50K. Expression of SM1/SM2/NM (%) in normal myometrium was 31.7/34.7/33.6 and 35.1/40.9/24 in hypertrophied myometrium. The increased SM2 and decreased NM expression in the hypertrophied state was statistically significant (P < 0.05). MHC isoform distribution in myomatous tissue was similar to normal myometrium (36.3/35.3/29.4). In vivo expression of MLC17a increased from 25.5% in normal to 44.2% in hypertrophied (P < 0.001) myometrium. Phosphorylation levels of MLC20 upon maximal Ca(2+)-calmodulin activation of skinned myometrial fibers were the same in normal and hypertrophied myometrial fibers. Maximal force of isometric contraction of skinned fibers (pCa 4.5, slack-length) was 2.85 mN/mm2 and 5.6 mN/mm2 in the normal and hypertrophied state, respectively (P < 0.001). Apparent maximal shortening velocity (Vmax(appt), extrapolated from the force-velocity relation) of myometrium rose from 0.13 muscle length s-1 (ML/s) in normal to 0.24 ML/s in hypertrophied fibers (P < 0.001).  相似文献   

15.
We have semi-synthesized 18 species of mixed chain phosphatidylethanolamines (PE) in which the sn-1 acyl chain is derived from stearic, arachidic, and behenic acids, and the sn-2 acyl chain is originated from cis,cis-octadecadienoic and cis, cis-eicosadienoic acids with the two methylene-interrupted double bonds located at various positions. These PEs constituting the bilayers in the aqueous dispersion were subjected to differential scanning calorimetric experiments. The Tm values associated with the gel-to-liquid crystalline phase transitions for these PEs are found to be significantly smaller than those of the saturated counterparts. Moreover, the magnitude of the Tm-lowering effect of acyl chain diunsaturation depends critically on the positions of the two methylene-interrupted cis double bonds in the sn-2 acyl chain. Specifically, if the sn-2 acyl chain is derived from cis, cis-octadecadienoic acid, the Tm-lowering effect has the following decreasing order: Delta9,12 > Delta6,9 > Delta12,15. For cis, cis-eicosadienoyl acyl chain, the Tm-lowering effect is stronger in the order of Delta10,13 > Delta11,14 > Delta8,11 > Delta5,8 > Delta14,17. Finally, a refined molecular model is presented that can adequately explain the Tm-lowering effect of sn-2 acyl chain diunsaturation. Moreover, this same refined molecular model can also be invoked to better interpret the Tm-lowering effect observed for sn-1 saturated/sn-2 monoenoic PE.  相似文献   

16.
Explores the problem of the sticky metaphor (SM) in psychoanalysis and psychoanalytic psychotherapy. SMs are recurring themes in the therapy that seem to reappear throughout the treatment and are often associated with a repetitive screen memory. Understanding the adhesiveness of SMs may provide important data on both transference and countertransference issues in therapy. An SM may often result from a complimentary identification in the therapist, which may or may not also be related to the analyst's experiencing similar conflicts in his or her own personality. A case of SM in therapy with a 34-yr-old White man illustrates the recurrence of a metaphor based on an early traumatic experience. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

17.
In a previous paper we described the experiments and the framework of a model for the exchange of monooleoylphosphatidylcholine with a single egg phosphatidylcholine membrane. In the present paper a model is presented that relates the experimentally measured apparent characteristics of the overall kinetics of lysolipid exchange to the true rates of lysolipid exchange and interbilayer transfer. It is shown that the adsorption of the lysolipid follows two pathways: one through the adsorption of lipid monomers and other through the fusion of micelles. The desorption of lysolipid follows a single pathway, namely, the desorption of monomers. The overall rate of fast desorption under convective flow conditions gives the true rate of monomer desorption from the outer membrane monolayer. The overall rate of both slow lysolipid uptake and slow desorption gives the rate of interbilayer transfer. Because of the uneven distribution of lysolipid between the two monolayers during its uptake, one of the membrane monolayers is apparently extended relative to the other. This relative extension of one of the monolayers induces a monolayer tension. The induced monolayer tension can increase up to 7 mN.m-1, when most of the intercalated lysolipid only partitions into the monolayer facing the lysolipid solution. This value is similar to the measured value for the critical monolayer tension of membrane failure, which is on the order of 5 mN.m-1. The similarity of the magnitudes of the induced monolayer tension during monooleoylphosphatidylcholine exchange and the monolayer tension of membrane failure suggests that the interbilayer lipid transfer may be affected by the formation of short living membrane defects. Furthermore, the pH-induced interbilayer exchange of phosphatidylglycerol is considered. In this case, it is shown that the rate of interbilayer transfer is a function of the phosphatidylglycerol concentration in the membrane.  相似文献   

18.
Pneumocystis carinii carinii and rat lung phospholipids contained 3-6% 1-alkyl-2-acyl glycerols composed of the glyceryl ether species, 1-O-octadecyl glycerol (batyl alcohol), 1-O-octadec-9-enyl glycerol (selachyl alcohol), 1-O-hexadecyl glycerol (chimyl alcohol), and 1-O-hexadec-9-enyl glycerol. Of the major phospholipid classes, phosphatidylinositol (PI) and phosphatidylserine contained the highest percentage of alkyl acyl glycerols. Methylprednisolone treatment caused an increase in alkyl acyl PI of rat lung lipids from 12% to 45%. As the PI concentration in lung phospholipids increases in rats treated with methylprednisolone, the increase in alkyl acyl PI was substantial; the proportions of alkyl acyl phosphatidylethanolamine and alkyl acyl lyso phosphatidylcholine (PC) also increased. Pneumocystis phospholipids contained higher proportions of alkyl acyl PC than the phospholipids of the lungs from normal and immunosuppressed uninfected rats. The glyceryl ether compositions of P. carinii carinii PC and lyso PC were similar, which suggests that lyso PC in the organism is derived by phospholipase A2 action on PC. This was not the case for PC and lyso PC of the lung controls. Analysis of the free fatty alcohols, precursors of glyceryl ethers identified only saturated species in P. carinii carinii and rat lung controls. Thus, the introduction of a double bond in the alcohol moiety of glyceryl ethers occurs after formation of the ether linkage between fatty alcohol and the glyceryl backbone.  相似文献   

19.
Based on the structural properties of cholesterol and egg phosphatidylcholine, and on the assumption that the van der Waals' type attactive interaction between the steroid nucleus and the fatty acyl chains provides a stabilizing force for the cholesterol-egg phosphatidylcholine complex, some specific orientation and configurations of the fatty acyl chains around the steroid nucleus in the interacting system are proposed in terms of an optimal packing. The proposed model suggests thathe saturated chains are largely facing the flattened (alpha) surface of the steroid nucleus of cholesterol, while the unsaturated chains can interact with both the alpha and beta surfaces of the steroid nucleus. It is also suggested that the angular methyl groups on the beta surface of the steroid nucleus lock the unsaturated fatty acyl chain in a relatively immobile configuration. Experimental evidence which provides support for the proposed stereochemical model is presented.  相似文献   

20.
BACKGROUND: Although fine-needle aspiration (FNA) is 90% sensitive in the detection of papillary carcinoma (PC) of the thyroid, its specificity has been reported as low as 52%. Consequently, patients who have an FNA suspicious for PC may undergo operation for a benign process. The ribonucleoprotein telomerase has been noted to be activated in a wide variety of carcinomas. We examined 30 PCs for telomerase activity to determine whether this would be a useful adjunct to FNA in the diagnosis of lesions suspicious for PC. METHODS: Standard telomere repeat amplification protocol assays were performed on fresh frozen tissue samples from 30 PCs, 3 benign nodules, and 10 normal thyroids. RESULTS: Telomerase activity was documented in 20 of 30 (67%) of the PCs, 0 of 3 benign nodules, and 0 of 10 normal thyroids. In all, 11 of the 20 PCs had FNA cytology that was nondiagnostic of PC, and 2 of the benign nodules had FNA that was suspicious for PC. CONCLUSIONS: The telomerase assay appears useful in the distinction of benign from malignant thyroid lesions that have FNA suspicious for but not diagnostic of PC. On the basis of these findings, a prospective trial examining telomerase activity in FNAs suspicious for thyroid cancer has been initiated.  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