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.
Interactions of phospholipid monolayers with carbohydrates   总被引:10,自引:0,他引:10  
Surface pressure studies of phospholipid monomolecular films of dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylcholine (DPPC) formed at an air/water interface have been made and the effects on the films studied when various carbohydrates are present in the subphase. The results obtained show that at a given temperature, the area per molecule of DPPC increases with increasing concentration of the carbohydrate in the subphase. The carbohydrate which has the greatest expanding effect on the phospholipid monolayer is glycerol, followed in turn by trehalose, sucrose, glucose, raffinose, and inositol. The mechanism of monolayer expansion by glycerol is different from that observed in other carbohydrates, as the following experiments demonstrate. Below the phase transition temperature of DPPC, the area per molecule of DPPC at a pressure of 12.5 dyn/cm is the same with and without glycerol in the subphase. However, when the monolayer is heated to a temperature above the phase transition temperature for DPPC, the area/molecule on glycerol is considerably greater than the area/molecule on water at the same surface pressure. Cooling the monolayer back to the lower temperature produces an area/molecule of DPPC which is identical on both water and glycerol subphases. Glycerol therefore has no effect on the low-temperature (condensed) monolayers but causes expansion of the high-temperature (expanded) monolayers. By contrast with glycerol, both trehalose and sucrose interact with the DPPC monolayer producing an increased area/molecule over that observed on water, both with low-temperature (condensed) monolayers and with the high-temperature (expanded) monolayers. The efficiency of these carbohydrates at expanding the monolayer films (with the exception of glycerol) shows a strong correlation with their ability to stabilize membrane structure and function at low water contents.  相似文献   

2.
1. The interactions between cytochrome c (native and [(14)C]carboxymethylated) and monolayers of phosphatidylcholine, phosphatidic acid and cardiolipin at the air/water interface was investigated by measurements of surface radioactivity, pressure and potential. 2. On a subphase of 10mm-or m-sodium chloride, penetration of cytochrome c into egg phosphatidylcholine monolayers, as measured by an increase of surface pressure, and the number of molecules penetrating, as judged by surface radioactivity, were inversely proportional to the initial pressure of the monolayer and became zero at 20dynes/cm. The constant of proportionality was increased when the cytochrome c was carboxymethylated or decreased when the phospholipid was hydrogenated, but the cut-off point remained at 20dynes/cm. 3. Penetrated cytochrome c could be removed almost entirely by compression of the phosphatidylcholine monolayer above 20dynes/cm. 4. With phosphatidic acid and cardiolipin monolayers on 10mm-sodium chloride the binding of cytochrome c was much stronger and cytochrome c penetrated into films nearing the collapse pressure (>40dynes/cm.). The penetration was partly electrostatically facilitated, since it was decreased by carrying out the reaction on a subphase of m-sodium chloride, and the relationship between the surface pressure increment and the initial film pressure moved nearer to that observed with phosphatidylcholine. 5. Surface radioactivity determinations showed that [(14)C]carboxymethylated cytochrome c was still adsorbed on phosphatidic acid and cardiolipin monolayers after the cessation of penetration. This adsorption was primarily electrostatic in nature because it could be prevented and substantially reversed by adding m-sodium chloride to the subphase and there was no similar adsorption on phosphatidylcholine films. 6. The penetration into and adsorption on the three phospholipid monolayers was examined as a function of the pH of the subphase and compared with the state of ionization of both the phospholipid and the protein, and the area occupied by the latter at an air/water interface. 7. It is concluded that the binding of cytochrome c to phospholipids can only be partially understood by a consideration of the ionic interaction between the components and that subtle conformational changes in the protein must affect the magnitude and stability of the complex. 8. If cytochrome c is associated with a phospholipid in mitochondria then cardiolipin would fulfil the characteristics of the binding most adequately.  相似文献   

3.
Steady-state and time-resolved fluorescence properties of the 7-nitrobenz-2-oxa-1, 3-diazole-4-yl (NBD) fluorophore attached either to the sn-2 acyl chain of various phospholipids (phosphatidylcholine, phosphatidylethanolamine, phosphatidylserine, and phosphatidic acid) or to the polar headgroup of phosphatidylethanolamine were studied after insertion of these NBD-labeled lipid probes into unilamellar vesicles of phosphatidylcholine, phosphatidylglycerol, phosphatidic acid, and phosphatidylserine. The fluorescence response of the NBD group was observed to strongly depend on the chemical structure and physical state of the host phospholipids and on the chemical structure of the lipid probe itself. Among the various fluorescence parameters studied, i.e., Stokes' shifts, lifetimes, and quantum yields, the quantum yields were by far the most affected by these structural and environmental factors, whereas the Stokes' shifts were practically unaffected. Thus, depending on the phospholipid probe and the host phospholipid, the fluorescence emission of the NBD group was found to vary by a factor of up to 5. Careful analysis of the data shows that for the various couples of probe and host lipid molecules studied, deexcitation of the fluorophore was dominated by nonradiative deactivation processes. This great sensitivity of the NBD group to environmental factors originates from its well-known solvatochromic properties, and comparison of these knr values with those obtained for n-propylamino-NBD in a set of organic solvents covering a large scale of polarity indicates that in phospholipids, the NBD fluorophore experiences a dielectric constant of around 27-41, corresponding to a medium of relatively high polarity. From these epsilon values and on the basis of models of the dielectric transition that characterizes any water-phospholipid interface, it can be inferred that for all of the phospholipid probes and host phospholipids tested, the NBD group is located in the region of the polar headgroups, near the phosphoglycerol moiety of the lipids.  相似文献   

4.
Lipids of a T Strain of Mycoplasma   总被引:10,自引:5,他引:5       下载免费PDF全文
Cholesterol, free fatty acids, and phosphatidic acid are the predominant lipids of a T strain of Mycoplasma. The remaining neutral lipids are composed of cholesteryl esters, triglycerides, and diglycerides. Three glucose-containing glycolipids are present in trace amounts. In addition to phosphatidic acid, the phospholipids are comprised of phosphatidyl glycerol, diphosphatidyl glycerol, and phosphatidyl ethanolamine. Another polar lipid was found to be ninhydrin-positive and phosphate-free. It appears to be a diamino hydroxy compound containing adjacent fatty acid ester and N-acyl groups.  相似文献   

5.
Phospholipids are the major components of pulmonary surfactant. Dipalmitoylphosphatidylcholine is believed to be especially essential for the surfactant function of reducing the surface tension at the air-liquid interface. Surfactant protein A (SP-A) with a reduced denatured molecular mass of 26-38 kDa, characterized by a collagen-like structure and N-linked glycosylation, interacts strongly with a mixture of surfactant-like phospholipids. In the present study the direct binding of SP-A to phospholipids on a thin layer chromatogram was visualized using 125I-SP-A as a probe, so that the phospholipid specificities of SP-A binding and the structural requirements of SP-A and phospholipids for the binding could be examined. Although 125I-SP-A bound phosphatidylcholine and sphingomyeline, it was especially strong in binding dipalmitoylphosphatidylcholine, but failed to bind phosphatidylglycerol, phosphatidylinositol, phosphatidylethanolamine, and phosphatidylserine. Labeled SP-A also exhibited strong binding to distearoylphosphatidylcholine, but weak binding to dimyristoyl-, 1-palmitoyl-2-linoleoyl-, and dilinoleoylphosphatidylcholine. Unlabeled SP-A readily competed with labeled SP-A for phospholipid binding. SP-A strongly bound dipalmitoylglycerol produced by phospholipase C treatment of dipalmitoylphosphatidylcholine, but not palmitic acid. This protein also failed to bind lysophosphatidylcholine produced by phospholipase A2 treatment of dipalmitoylphosphatidylcholine. 125I-SP-A shows almost no binding to dipalmitoylphosphatidylglycerol and dipalmitoylphosphatidylethanolamine. The addition of 10 mM EGTA into the binding buffer reduced much of the 125I-SP-A binding to phospholipids. Excess deglycosylated SP-A competed with labeled SP-A for binding to dipalmitoylphosphatidylcholine, but the excess collagenase-resistant fragment of SP-A failed. From these data we conclude that 1) SP-A specifically and strongly binds dipalmitoylphosphatidylcholine, 2) SP-A binds the nonpolar group of phospholipids, 3) the second positioned palmitate is involved in dipalmitoylphosphatidylcholine binding, and 4) the specificities of polar groups of dipalmitoylglycerophospholipids also appear to be important for SP-A binding, 5) the phospholipid binding activity of SP-A is dependent upon calcium ions and the integrity of the collagenous domain of SP-A, but not on the oligosaccharide moiety of SP-A. SP-A may play an important role in the regulation of recycling and intra- and extracellular movement of dipalmitoylphosphatidylcholine.  相似文献   

6.
Phase transitions in bilayers and monolayers of various synthetic phospholipids with different chain lengths as well as different polar head groups were studied by differential scanning calorimetry or with the film balance technique, respectively. With the film balance, area versus temperature curves (isobars) were recorded at different surface pressures. The monolayer phase transition from the fluid-condensed to the fluid-expanded phase is shifted towards higher temperature when the lateral pressure in the monolayer is increased. The temperature dependence of the equilibrium pressure as well as the magnitude of the area change at the transition depends only on the nature of the phospholipid head group and not on the chain length of the hydrocarbon chains of the lipid. Phospholipids with strong intermolecular attractive interactions between the head groups show low values for dpi/dTm and for the area change, deltaf, whereas phospholipids with negatively charged head groups without intermolecular attractive forces exhibit higher values for dpi/dTm and deltaf. The shift of the monolayer phase transition temperature when increasing the chain length of the lipid is almost identical to the shift in Tm observed for the bilayer system of the same phospholipids. A comparison of monolayer and bilayer systems on the basis of the absolute value of the molecular area of the phospholipid in the bilayer gel phase and the change in area at the bilayer and monolayer transition leads to the following conclusions. The behaviour of the bilayer system is very similar to that of the respective monolayer system at a lateral pressure of approx. 30 dyne/cm, because at this pressure the absolute area and the area change in both systems are the same. Further support for this conclusion comes from the experimental finding that a lateral pressure of 30 dyne/cm the shift in Tm due to the increase in charge when the methyl ester of phosphatidic acid is investigated is the same for the bilayer and the monolayer system.  相似文献   

7.
The surface properties at the air/water interface and the interaction of branched chain polymeric polypeptides with a general formula poly[Lys-(DL -Alam-X1)], where X = Π (AK), Ser (SAK), or Glu (EAK), with phospholipids were investigated. Polylysine derivatives with polycationic (SAK, AK) or amphoteric (EAK) were capable to spread and form stable monomolecular layers. The stability of monolayers at the air/water interface was dependent on the side-chain terminal amino acid residue of polymers and can be described by SAK < AK < EAK order. The area per amino acid residue values calculated from compression isotherms were in the same range as compared to those of linear poly-α-amino acids and proteins. Moreover, these polymers interact with phospholipid monomolecular layers composed of dipalmitoyl phosphatidyl choline (DPPC) or DPPC/PG (PG: phosphatidyl glycerol; 95/5, mol/mol). Data obtained from compression isotherms of phospholipids spread on aqueous polymer solutions at different initial surface pressure indicated that insertion into lipid monolayers for SAK or AK is more pronounced than for EAK. The interaction between branched polypeptides and phospholipid membranes was further investigated using lipid bilayers with DPPC/PG and fluorescent probes located either at the polar surface [1-(4-trimethylammonium-phenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) sodium anilino naphthalene sulfonate (ANS)] or within the hydrophobic core (DPH) of the liposome. Changes in fluorescence intensity and in polarization were observed when TMA-DPH or ANS, but not DPH were used. Comparative data also indicate that all three polymers interact only with the outer surface of the bilayer, but even the most marked penetration of polycationic polypeptide (SAK) did not result in alteration of the ordered state of the alkyl chains in the bilayer. Taken together, data obtained from mono- or bilayer experiments suggest that the interaction between branched polymers and phospholipids are highly dependent on the charge properties (Ser vs Glu) and on the identity (Ser vs Ala) of side-chain terminating amino acids. The binding of polymers to the model membranes could be mainly driven by electrostatic forces, but the significant role of hydrophilic properties in case of SAK cannot be excluded. © 1998 John Wiley & Sons, Inc. Biopoly 46: 169–179, 1998  相似文献   

8.
Different phospholipids, except the choline-containing phospholipids phosphatidylcholine, lysophosphatidylcholine, and sphingomyelin, formed complexes with the dye Victoria blue R, which selectively partitioned into the chloroform phase of chloroform/ethylene glycol/glycerol biphasic solvent system, and were quantitatively estimated at 590 nm. Considerable amounts of water, alcohols, nonlipid phosphates, neutral lipids, free fatty acids, and some detergents did not interfere with the formation of phospholipid-dye complexes. This special advantage of the method described allowed combined phospholipid extraction and estimation procedures in one test tube. Because of its high sensitivity (about 24.00 OD units/mumol of phosphatidic acid and about 10.25 OD units/mumol of other phospholipids), specificity, and simplicity, the proposed phospholipid assay appears to be very useful for rapid analyses of lipid extracts as well as TLC spots or suspensions of biological materials, as demonstrated for membranes and cells of Micrococcus lysodeikticus. The applicability of the dye Victoria blue B to the quantitative determination of phospholipids, except phosphatidylcholine, lysophosphatidylcholine, and sphingomyelin, at 605 nm using chloroform/ethylene glycol/glycerol/water and pentane (hexane)/ethyl acetate/isopropanol/water biphasic solvent systems with similar sensitivities and of sodium dodecyl sulfate in the pentane-containing system with high sensitivity (22.96 OD units/mumol) is also shown. The adaptation of this phospholipid assay to the determination of phospholipases C and D and to the differential quantitation of choline-containing phospholipids using additional phospholipid estimation techniques is discussed.  相似文献   

9.
Phospholipid flip-out controls the cell cycle of Escherichia coli   总被引:2,自引:0,他引:2  
Phospholipids are the principal constituents of biological membranes. In Escherichia coli, phospholipids are involved in the metabolism of other envelope constituents such as lipoprotein, lipopolysaccharide, certain envelope proteins and peptidoglycan. They are also involved in the regulation of the cell cycle. DNAA, the key protein in the initiation of chromosome replication, is activated by acidic phospholipids only when these are in fluid bilayers, whilst interruptions of phospholipid synthesis inhibit both the initiation of chromosome replication and cell division. The transmembrane movement or flip-flop of phospholipids from one monolayer to the other requires the passage of the polar head group through the hydrophobic core of the bilayer. Hence, in many systems, flip-flop is a slow process with half-time of days. Flip-flop accompanies the formation of non-bilayer structure. Such structures form under certain conditions of packing density and composition and have been observed both in vitro and in vivo. In bacteria, flip-flop appears to be extremely rapid, with half-times as fast as 3 min being observed. However, such rapid flip-flop may not be characteristic of all phospholipids. The asymmetrical distribution of phosphatidylethanolamine in the plasma membrane of Bacillus megaterium has been attributed to the existence of two classes of this phospholipid. In E. coli, studies of the metabolic turnover of phosphatidylserine, phosphatidylglycerol and phosphatidic acid also reveal the existence of distinct classes of these phospholipids. In this article I propose that, in E. coli, a class of phospholipids does indeed escape the rapid flip-flop mechanism; this class probably includes a subpopulation of the acidic phospholipids. Therefore during the cell cycle these phospholipids accumulate in the inner monolayer of the cytoplasmic membrane and so cause an increase in its packing density; at a critical density, phospholipids "flip out" from the inner to the outer monolayer. This flip-out occurs once per cycle and initiates cell cycle events.  相似文献   

10.
Relation between cellular phospholipids and L-glutamic acid excretion was investigated using Corynebacterium alkanolyticum GL–21 (a glycerol auxotroph).

When strain GL–21 was cultured in glycerol-limited medium which contained n-hexadecane, acetic acid or fructose as carbon source, there occurred the limitation of cellular phospholipid content and the over-accumulation of L-glutamic acid in the broth. Two-dimensional thin-layer chromatograms provided evidence that both the parent and the mutant strains contained the same phospholipids such as cardiolipin, phosphatidylethanolamine, phosphadityl-glycerol, phosphatidylinositol and phosphatidic acid. Limited supply of glycerol to the mutant did not greatly alter the proportions of the individual phospholipids.  相似文献   

11.
Adsorption of chloroform, halothane, enflurane and diethyl ether on the air/water interface was compared with adsorption on the dipalmitoylphosphatidylcholine monolayer, spread on the air/water interface, at four compressional states; 88.5, 77.0, 66.5 and 50.5 A2 surface area per phosphatidylcholine molecule. Anesthetics were administered from the gas phase. The affinities of these agents to the phosphatidylcholine monolayer varied according to the state of the monolayer. Chloroform and halothane showed a stronger affinity to the highly compressed phosphatidylcholine monolayer (50.5 A2) than to the expanded monolayer (88.5 A2) or to the air/water interface without the monolayer. Diethyl ether behaved in reverse; a stronger affinity to the expanded monolayer was exhibited than to the compressed monolayer. Enflurane showed the highest affinity to the intermediately compressed monolayer (77.0 A2). The adsorption isotherm of anesthetics to the monolayer was characterized by atypical Langmuir-type, in which available number of binding sites changed when anesthetics were adsorbed. The mode of adsorption onto the monolayer was dissimilar to adsorption onto air/water interface, where adsorption followed the Gibbs surface excess. A theory is presented to explain the above differences. The adsorbed anesthetic molecules do not stick to phosphatidylcholine molecules but penetrate into the monolayer lattice and occupy the phosphatidylcholine sites at the interface. Quantitative agreement between the theory and the experimental data was excellent. For the monolayer at 50.5 A2 compression, the changes in the transfer free energy accompanying the anesthetic adsorption from the gas phase to the monolayer were in the order of chloroform greater than halothane greater than enflurane greater than diethyl ether, in agreement with the clinical potencies.  相似文献   

12.
The phospholipid composition of the butyric acid-producing clostridia is responsive to the degree of enrichment of the lipids with cis-unsaturated fatty acids. When Clostridium butyricum and Clostridium beijerinckii are grown on oleic acid in media devoid of biotin, the acyl and alk-1-enyl chains of the phospholipids become highly enriched with 18:1 and C19-cyclopropane. Under these conditions there is a marked increase in the glycerol acetals of the major plasmalogens of these organisms. We have grown both species on mixtures of palmitate and oleate in the absence of biotin. The alk-1-enyl chains were highly enriched with C18-unsaturated and C19-cyclopropane residues at all but the highest ratios of palmitate to oleate (80:20, w/w) added to the medium. At ratios of palmitate to oleate greater than or equal to 40:60, the saturated acid was incorporated predominantly into the phospholipid acyl chains in both organisms. The effects of increasing unsaturation of the acyl chains as the ratio of oleate to palmitate was increased was examined in C. butyricum. In cells grown on mixtures of palmitate and oleate equal to or exceeding 40% palmitate, the ratio of glycerol acetal lipid to total phosphatidylethanolamine (PE) was relatively constant. As the proportion of oleic acid added to the medium was increased, the ratio of glycerol acetal lipid to PE increased from 0.7 to 2.0. Thus the ratio of the polar lipids appears to respond to the content of phospholipids that contain two unsaturated chains. The fraction of PE present as plasmalogen remained relatively stable (0.82 +/- 0.05) at varying ratios of medium oleic and palmitic acids. Both the glycerol acetal of ethanolamine plasmalogen, and ethanolamine plasmalogen, are shown to be 80% or more in the outer monolayer of the cell membrane. These two polar lipids represent approx. 50% of the phospholipids in cells grown on exogenous fatty acid. The bulk of the remainder is polyglycerol phosphatides. We suggest that the ability of both species to grow with highly unsaturated membranes is related to their ability to modulate their polar lipid composition.  相似文献   

13.
We studied the interactions of tetracycline antibiotics, TCs, with phospholipid monolayers with the two-fold aim of elucidating the mechanism of action of TCs and to provide a first step for the realization of bio-mimetic sensor for such drugs by means of the Langmuir-Blodgett technique. Preliminary surface tension studies demonstrated that surface activity of tetracycline is moderate and dependent on the pH of the subphase. We selected three phospholipids having hydrophobic chains of the same length but differing in the polar head structures, i.e. dipalmitoylphosphatidylcholine, dipalmitoylphosphatidylethanolamine, and dipalmitoylphosphatidic acid. Surface pressure- and surface potential- area isotherms were employed to investigate the behavior of the phospholipid monolayers at the water-air interface when tetracycline was added to the aqueous subphase. Analysis of the results indicated that the electrostatic interaction is the driving force for migration of tetracycline towards the interface where localized adsorption to the head groups occurs. Nevertheless, such interactions appear to be insufficient to promote penetration of tetracycline through the hydrophobic layer.  相似文献   

14.
Saponification of extracted tissue phospholipids yields a set of isolated glycerol 3-phosphoryl phospholipid polar headgroups from which semi-quantitative 31P NMR spectra can be obtained. The resonance signals from these molecules, which frequently have been reported as uncharacterized phosphate signals observed in perchloric acid extracts of tissue, can be used as an aid in the characterization of isolated phospholipids and of tissue phospholipid 31P NMR profiles. 31P NMR chemical-shift values of the resonances at pH 7 in water and relative to 85% phosphoric acid are: glycerol 3-phosphocholine (-0.13 delta), glycerol 3-phosphoethanolamine (0.42 delta), glycerol 3-phospho(monomethyl)ethanolamine (0.29 delta), glycerol 3-phospho(dimethyl)ethanolamine (0.16 delta), glycerol 3-phosphoserine (0.14 delta), glycerol 3-phosphoinositol (-0.07 delta), glycerol 3-phosphoglycerol (0.92 delta), bis(glycerol 3-phospho)glycerol (0.79 delta), serine ethanolamine phosphodiester (-0.46 delta), glycerol 3-phosphate (0.60 delta; 4.29 delta at pH 10) glycerol 2-phosphate (0.15 delta; 3.92 delta at pH 10). In addition, analysis of extracted cancer tissue phospholipid samples yielded a new and uncharacterized polar headgroup fragment with a chemical-shift value of 0.29 delta that is independent of sample pH.  相似文献   

15.
A fluorescent indicator for the determination of pH in the vicinity of water/lipid interfaces was produced by the covalent linkage of fluoresceinisothiocyanate to Escherichia coli phosphatidylethanolamine. When embedded in monolayers spread on an air/water interface, its apparent pK was shown to be only slightly affected by the nature of the polar headgroups or the packing density of the host phospholipids. Its properties were not affected by the ion content of the subphase. For small unilamellar vesicles, its properties were only affected when localized in the inner layer. This probe could therefore be of value in the study of proton fluxes along biological membranes.  相似文献   

16.
This paper is part of a systematic study of the interactions of tetracycline antibiotics with phospholipid monolayers at the water-air interface. Tetracyclines are widespread antibiotics that undergo a series of protonation equilibria in solution, depending on the pH. The surface activity of tetracyclines was determined by means of surface tension measurements for three different systems, i.e. water, TRIS and McIlvaine-EDTA buffer. Surface pressure-molecular area and surface potential-molecular area isotherms were acquired for dipalmitoylphosphatidic acid monolayers on TRIS buffer (pH=7.0) and McIlvaine-EDTA buffer (pH=4.0) solution as a function of tetracycline concentration in the subphase. Comparative analysis of surface potential data, with the molecular dipole moment of tetracycline obtained from semiempirical calculations, provided information on the orientation of tetracycline at the interface. Surface pressure measurements as a function of monolayer compression were described, applying either a continuous partition model or Langmuir adsorption isotherms. The results obtained in the case of buffer solutions were compared to those obtained for tetracycline in water subphases. The analysis of the results indicated that electrostatic interactions dictate the migration of tetracycline to the monolayer interface. Penetration of the molecule to the lipophilic portion of the monolayer was unlikely, especially at high surface pressures. The results showed that stronger interactions are established between the zwitterionic tetracycline and the deprotonated phosphatidic group in TRIS buffer solution; in this case, tetracycline binds at the monolayer interface following a Langmuir type adsorption. In the case of water, where the monodeprotonated acid and the tetracycline zwitterions are the only species involved, the data can be described by continuous partition of tetracycline between interfacial and bulk phases. The same holds for McIlvaine-EDTA buffer subphases, although the high concentrations of citrate ions in this buffer competitively interfere with tetracycline association at the monolayer interface.  相似文献   

17.
The effects of pulmonary surfactant protein SP-B on the properties of monolayers of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylglycerol (DPPG), and a mixture of DPPC:DPPG (7:3, mol:mol) were studied using spread films at the air-water interface. The addition of SP-B to the phospholipid monolayers gave positive deviations from additivity of the mean areas in the films. At low protein concentrations (less than 45% amino acid residues which corresponds to 0.5 mol% or 10 weight% SP-B) monolayers of SP-B/DPPC, SP-B/DPPG and SP-B/(DPPC:DPPG) collapsed at surface pressures of about 70 mN.m-1, comparable to those of the lipids alone. At higher concentrations of SP-B in the protein-lipid monolayers, kink points appeared in the isotherms at about 40-45 mN.m-1, implying possible exclusion of material from the films, hence, changes in the original monolayer compositions. Calculated analyses of the monolayer compositions as a function of surface pressure indicated that nearly pure SP-B, associated with small amounts of phospholipid (2-3 lipid molecules per SP-B dimer), was lost from SP-B/DPPC, SP-B/DPPG, and SP-B/(DPPC:DPPG) films at surface pressures higher than 40-45 mN.m-1. The results are consistent with a low effectiveness of SP-B in removing saturated phospholipids, DPPC or DPPG, from the spread SP-B/phospholipid films.  相似文献   

18.
Three neoglycolipids 1a–c derived from glycerol were synthezised and their molecular arrangements were studied at the air/water interface through Langmuir–Blodgett technique. The common structural features of these neoglycolipids are: a triethyleneglycol spacer at C-2 of glycerol, a GlcNAc hydrophilic head group at the end of the spacer, two saturated aliphatic chains at C-1 and C-3 of glycerol, linked by ether bonds. Compounds 1a–c differ only by the length of their lipid moieties. By increasing the hydrocarbon chain length from C11 to C16, a densely packed state was reached in the monolayer. The compression isotherms displayed an expanded state during the whole compression for compounds 1a and 1c bearing two C11 or one C11 and one C16 chains. Compound 1b bearing two C16 chains displayed a quite different interfacial behavior. The transfer of these monolayers onto a solid substrate can be obtained only with a trigger molecule, a phosphatidic acid.  相似文献   

19.
The fatty acid composition of wheat seedling roots changed in response to temperature. As temperature declined, the level of linolenic acid increased and the level of linoleic acid decreased. The distribution of phospholipid classes was not influenced by temperature. Phosphatidyl choline and phosphatidyl ethanolamine were the predominant phospholipids isolated and comprised 85% of the total lipid phosphorus. Smaller quantities of phosphatidyl glycerol, phosphatidyl inositol, phosphatidic acid, and phosphatidyl serine were isolated. The fatty acid composition of phosphatidyl choline and phosphatidyl ethanolamine were the same and temperature affected the fatty acid composition of both phospholipids in the same manner.Growth in the presence of the substituted pyridazinone, BASF 13 338 (4-chloro-5-dimethylamino-2-phenyl-3(2H)pyridazinone), reduced the level of linolenic acid and increased the level of linoleic acid in the phosphatidyl choline, phosphatidyl ethanolamine, and total polar lipid fractions. BASF 13 338 did not affect the levels of palmitate, stearate, and oleate or the distribution of phospholipid classes.Respiration rates of wheat root tips were measured over a range of temperatures. The respiration rate declined as the temperature decreased. Neither the temperature at which the tissue was grown nor BASF 13 338 treatment influenced the ability of root tips to respire at any temperature from 4 to 30 C. The results indicated that the relative proportion of linolenic acid to linoleic acid did not influence the plants ability to grow and respire over the range of temperatures tested.  相似文献   

20.
In the present work, it is shown that the photobleaching technique as well as experimentation based on fluorescence recovery after bleaching can be extended to monolayers spread at the air-water interface. A mathematical model is derived which allows the determination of translational diffusion coefficients of species diffusing in such a system. Using 12-(9-anthroyl)stearic acid (anthroylstearate) as a fluorescent probe, dispersed either in dipalmitoylphosphatidylcholine or in dipalmitoylphosphatidylglycerol in various conditions of subphase ionic composition and surface pressure of the monolayer, including phase transition domains, we are led to the following conclusions: 1. Anthroylstearate molecules seem to aggregate in 'microdomains' where their fluorescence properties remain unchanged regardless of the compression states of the host monolayer. 2. In any case, a break in the diffusion constants appears on compressing films of both dipalmitoylphosphatidylcholine and dipalmitoylphosphatidylglycerol. In particular, this break coincides with the liquid expanded to gel phase transition of these lipids when it occurs. 3. Diffusion of anthroylstearate in dipalmitoylphosphatidylglycerol depends strongly on the subphase ionic strength and on the nature of cations: Na+, Mg2+, Ca2+.  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