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 31 毫秒
1.
The effect of 2,3-seco-5 alpha-cholestan-2,3-dioic acid on the bilayer to hexagonal phase transition temperature of dielaidoylphosphatidylethanolamine is markedly dependent on pH. Above pH 6.56, the 2,3-seco-5 alpha-cholestan-2,3-dioic acid raises the temperature of this transition, i.e., it stabilizes the bilayer phase. At pH 6.56 there is little effect of this sterol derivative on the bilayer to hexagonal phase transition temperature of dielaidoylphosphatidylethanolamine. However, below pH 6.56, the 2,3-seco-5 alpha-cholestan-2,3-dioic acid markedly lowers the temperature of this transition. The promotion of hexagonal phase formation increases both with increasing mol fraction of this sterol derivative and with lower pH, particularly in the range between pH 6.56 and pH 5.0. Below about pH 6, 2,3-seco-5 alpha-cholestan-2,3-dioic acid also induces vesicle fusion as measured both by lipid mixing as well as by mixing of aqueous contents. For these assays vesicles made of phosphatidylethanolamine (made from egg phosphatidylcholine) and extruded through 0.2 micron pore membranes were used. At higher concentrations or at lower pH the 2,3-seco-5 alpha-cholestan-2,3-dioic acid induces some leakage of the contents of these vesicles. Nevertheless, with vesicles containing only 2 weight% sterol derivative, it was possible to demonstrate substantial mixing of aqueous contents of the vesicles over the pH range 3.5 to 5.5. Several of the properties of 2,3-seco-5 alpha-cholestan-2,3-dioic acid indicate that this compound may be useful in sensitizing vesicles to acid-induced fusion for the purpose of endocytic drug delivery.  相似文献   

2.
Dimethylsuberimidate was reacted with aqueous dispersions of dipalmitoylphosphatidylethanolamine, dimyristoylphosphatidylethanolamine, dilauroylphosphatidylethanolamine, and dielaidoylphosphatidylethanolamine at pH 10 and at pH 8. The amount of amidine dimer formation was about four times greater above the gel-to-fluid phase transition of each lipid than below the transition. The transition temperature of each phosphatidylethanolamine, measured by steady-state fluorescence anisotropy of cis-parinaric acid, was lower at pH 10 than at pH 8 or in water. The ability of dimethylsuberimidate to discriminate between phosphatidylethanolamines in the fluid and gel phases should allow use of this reagent to identify phosphatidylethanolamine species within the gel or fluid lipid phase.  相似文献   

3.
J J Cheetham  E Wachtel  D Bach  R M Epand 《Biochemistry》1989,28(22):8928-8934
The phase behavior of mixtures of cholesterol or epicholesterol with phosphatidylethanolamine was studied by differential scanning calorimetry and by X-ray diffraction. Discrete domains of cholesterol are detected by X-ray diffraction in the L alpha phase of phosphatidylethanolamine from egg yolk and synthetic dielaidoylphosphatidylethanolamine beginning at mole fractions of 0.35-0.4 cholesterol. Separate domains of crystalline epicholesterol can also be detected in the L alpha phase of dielaidoylphosphatidylethanolamine by X-ray diffraction at as little as 0.16 mole fraction of epicholesterol. This is a result of poor miscibility of the epicholesterol with dielaidoylphosphatidylethanolamine. Epicholesterol does not alter the L beta----L alpha transition or bilayer spacing. Epicholesterol also has little effect on the diameter of the cylinders in the hexagonal phase. Formation of the inverted hexagonal phase is facilitated by addition of small amounts of cholesterol (mole fraction less than 0.2) in both egg phosphatidylethanolamine and dielaidoylphosphatidylethanolamine. However, at higher mole fractions of cholesterol, the stability of the liquid-crystalline phase is found to increase markedly for dielaidoylphosphatidylethanolamine but not for egg phosphatidylethanolamine, indicating the importance of the structure of the acyl chains in controlling the relative stability of the lamellar and nonlamellar phases in these systems. In contrast to cholesterol, epicholesterol markedly lowers the L alpha----HII phase transition temperature at low mole fraction of sterol. This result demonstrates the importance of the orientation and motional properties of an additive in determining the L alpha----HII transition temperature.  相似文献   

4.
The interaction of all-trans-retinoic acid and all-trans-retinol with dielaidoylphosphatidylethanolamine has been studied by differential scanning calorimetry and 31P-NMR spectroscopy. Increasing concentrations of all-trans-retinoic acid up to a mol fraction of 0.09 were found to induce shifts to lower temperatures of both the L beta to L alpha and L alpha to hexagonal-HII phase transitions, with a slight decrease in the enthalpy change of the transitions. At higher concentrations no further effects on the transitions were observed, and this is interpreted as indicative of a limited miscibility of retinoic acid with the phospholipid. 31P-NMR spectroscopy confirmed that the L alpha to hexagonal-HII phase transition was shifted to lower temperatures in the presence of retinoic acid. On the other hand increasing concentrations of all-trans-retinol up to a mol fraction of 0.166, induced a progressive shift of the L beta to L alpha and the L alpha to hexagonal-HII phase transitions to lower temperatures. At higher concentrations the main gel to liquid-crystalline phase transition was further displaced to lower temperatures and the lamellar to hexagonal-HII phase transition was not observed in the thermograms. 31P-NMR spectroscopy indicated that retinol was able of inducing the phospholipid to adopt the hexagonal-HII phase at temperatures even below the main gel to liquid-crystalline phase transition temperature of the pure phospholipid.  相似文献   

5.
The phase behaviour, particularly the fluidity within each phase state and the transitions between them, of lipopolysaccharides and of their lipid moiety, free lipid A, of various species of Gram-negative bacteria, especially of Salmonella minnesota and Escherichia coli, has been investigated by applying mainly Fourier-transform infrared spectroscopy and differential scanning calorimetry. For enterobacterial strains, the transition temperatures of the gel----liquid crystalline (beta----alpha) phase transition of the hydrocarbon chains in dependence on the length of the sugar moiety are highest for free lipids A (around 45 degrees C) and lowest for deep rough mutant lipopolysaccharides (around 30 degrees C). Evaluating certain infrared active vibration bands of the hydrocarbon moiety, mainly the symmetric stretching vibration of the methylene groups around 2850 cm-1, it was found that, in the gel state, the acyl chains of lipopolysaccharides and free lipid A have a higher fluidity as compared with saturated and the same fluidity as compared with unsaturated phospholipids. This 'partial fluidization' of lipopolysaccharide below the transition temperature correlates with its reduced enthalpy change at that temperature compared to phospholipids with the same chain length. The fluidity depends strongly on ambient conditions, i.e. on the Mg2+ and H+ content: higher Mg2+ concentrations and low pH values make the acyl chains of free lipid A and lipopolysaccharide preparations significantly more rigid and also partially increase the transition temperature. The influence of Mg2+ is highest for free lipid A and decreases with increasing length of the sugar side chain within the lipopolysaccharide molecules, whereas the effect of a low pH is similar for all preparations. At basic pH, a fluidization of the lipopolysaccharide and lipid A acyl chains and a decrease in transition temperature take place. Free lipid A and all investigated rough mutant lipopolysaccharides exhibit an extremely strong lyotropic behaviour in the beta----alpha melting enthalpy but not in the value of the transition temperature. The phase transition is distinctly expressed only at water concentrations higher than 50-60%. A further increase of the water content still leads to an increase in the phase-transition enthalpy, particularly for lipopolysaccharides with a more complete sugar moiety. The fluidity of the hydrocarbon chains is shown to be an important parameter with respect to the expression of biological activities.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
Differential scanning calorimetry (DSC) has been employed to determine the effect of five commonly employed extrinsic potential-sensitive probes on phase transitions of multilamellar suspensions of L-alpha-dimyristoylphosphatidylcholine (DMPC). At mol% values of less than five, the effect of these probes on the excess heat capacity curve in the vicinity of the gel to liquid crystal phase transition can be described by an equation based on the formation of ideal solutions in both phases. Even at up to 4 mol%, these dyes only moderately reduce the enthalpy change associated with this transition, but cause a marked decrease in the size of the cooperative unit parameter. The excess heat capacity profile for diS-C3-(5) is represented by the ideal solution equation, even at 12 mol%, whereas the suspensions with the other probes present at this level have profiles covering large temperature ranges. Multiple peaks appear at the higher levels for the negatively charged oxonols V and VI, and merocyanine 540, a result consistent with the presence of well-defined microdomains or even phase separation. The enthalpy change associated with the transition near 15 degrees C involving packing in the headgroup region is decreased significantly, indicating that the probes probably affect the lipid headgroup conformation, even at low levels. The cyanine probe diS-C3-(5) causes the heat capacity profile of small unilamellar vesicles to be transformed very rapidly into one similar to that of the vortexed lipid preparations, presumably by a dye-mediated vesicle fusion process, enhanced by the surface location of this probe. All our results are consistent with diS-C3-(5) being located on the surface of the bilayer in both phases, but a penetration of the other probes into the hydrocarbon region, at least in the liquid crystal phase.  相似文献   

7.
R M Epand 《Biochemistry》1985,24(25):7092-7095
The bilayer to hexagonal phase transition temperatures of dielaidoylphosphatidylethanolamine and 1-palmitoyl-2-oleoylphosphatidylethanolamine are 65.6 and 71.4 degrees C, respectively. Using high-sensitivity differential scanning calorimetry, I have shown that these transition temperatures are extremely sensitive to the presence of small amounts of other lipid components. For example, at a mole fraction of only 0.01, dilinolenin lowers the bilayer to hexagonal phase transition temperature of 1-palmitoyl-2-oleoyl-phosphatidylethanolamine by 8.5 degrees C. Other diacylglycerols have similar effects on this transition temperature, although the degree of unsaturation of the acyl chains has some effect, with distearin being less potent. In comparison, the 20-carbon alkane eicosane lowers this transition temperature by 5 degrees C, while palmitoyl-lysolecithin raises it by 2.5 degrees C. Similar effects of these additives on the bilayer to to hexagonal phase transition temperature are observed with dielaidoylphosphatidylethanolamine. At these concentrations of additive, there is no effect on the gel-state to liquid-crystalline-state transition temperature. The observed shifts in the temperature of the bilayer to the hexagonal phase transition can be qualitatively interpreted in terms of the effects of these additives on the hydrophilic surface area and on the hydrophobic volume. Substances expanding the hydrophobic domain promote hexagonal phase formation and lower the bilayer to hexagonal phase transition temperature. The sensitivity of the bilayer to hexagonal phase transition temperature to the presence of additives is at least as great as that which has been observed for any other lipid phase transition.  相似文献   

8.
Abstract

Liposomes made of dipalmitoylphosphatidylcholine (DPPC2), dipalmitoyl-phosphatidylglycerol (DPPG), and different long-chain fatty alcohols were investigated with respect to their colloidal stability, chain-melting phase transition temperature, and temperature dependent inter-vesicle fusion. In particular, the practical usefulness of the stoichiometric 1/2 (mol/mol) mixtures of the phospholipids and fatty alcohols, mainly elaidoyl alcohol (EL-OH) were studied. The mole fraction of DPPG in the bilayers of such vesicles affects crucially the colloidal stability of the resulting lipid suspensions; at least 15 mol-% of DPPG (relative to DPPC) must be incorporated into the bilayers in order to make the liposome suspension colloidally sufficiently stable at room temperature. The corresponding DPPC/DPPG/EL-OH (0.85/0.15/2) mixed lipid vesicles undergo a lamellar-gel to inverted hexagonal (HIT) phase transition at 52.7°C, however, and then fuse and aggregate massively. The related phase transition temperature of the DPPC/DPPG/palmitelaidoyl alcohol (0.85/0.15/2) mixture is 48.4°C. This indicates that the chain-melting phase transition temperature of the investigated lipid mixtures is rather sensitive to the alcohol chain-length. This transition temperature is independent, however, of the bulk proton concentration in the pH region between 4.9 and 7.2. Stoichiometric 1/2 mixtures of phospholipids and EL-OH have a high propensity for the inter-vesicle fusion at 42°C and neutral pH. The reason for such fusion 10°C below the lamellar-to-nonlamellar phase transition temperature are the defects that are generated during the chain-melting of the (partly segregated) phospholipid component at 42°C; the proximity of the lamellar to non-lamellar phase transition temperature of the phospholipid/fatty alcohol (1/2) complex at 52°C also plays an important role.  相似文献   

9.
A number of sugars lowered the bilayer to hexagonal phase transition temperature of dielaidoylphosphatidylethanolamine. Disaccharides had the greatest effect followed by sugar alcohols. The monosaccharides, glucose and galactose had no effect on this phase transition temperature. The sugars promoted vesicle leakage only under conditions where the lipid was near its hexagonal phase transition temperature. Leakage from lipids in the bilayer state was inhibited by the sugars. Polyols, such as sorbitol, promote hexagonal phase formation and alter membrane permeability. These membrane effects may contribute to the damage caused by sorbitol accumulation in certain tissues of diabetic patients.  相似文献   

10.
Differential scanning calorimetry (DSC) and pressure perturbation calorimetry (PPC) were used to characterize thermal phase transitions, membrane packing, and volumetric properties in multilamellar vesicles (MLVs) composed of the polar lipid fraction E (PLFE) isolated from the thermoacidophilic archaeon Sulfolobus acidocaldarius grown at different temperatures. For PLFE MLVs derived from cells grown at 78 degrees C, the first DSC heating scan exhibits an endothermic transition at 46.7 degrees C, a small hump near 60 degrees C, and a broad exothermic transition at 78.5 degrees C, whereas the PPC scan reveals two transitions at approximately 45 degrees C and 60 degrees C. The endothermic peak at 46.7 degrees C is attributed to a lamellar-to-lamellar phase transition and has an unusually low DeltaH (3.5 kJ/mol) and DeltaV/V (0.1%) value, as compared to those for the main phase transitions of saturated diacyl monopolar diester lipids. This result may arise from the restricted trans-gauche conformational changes in the dibiphytanyl chain due to the presence of cyclopentane rings and branched methyl groups and due to the spanning of the lipid molecules over the whole membrane. The exothermic peak at 78.5 degrees C probably corresponds to a lamellar-to-cubic phase transition and exhibits a large and negative DeltaH value (-23.2 kJ/mol), which is uncommon for normal lamellar-to-cubic phospholipid phase transformations. This exothermic transition disappears in the subsequent heating scans and thus may involve a metastable phase, which is irreversible at the scan rate used. Further, there is no distinct peak in the plot of the thermal expansion coefficient alpha versus temperature near 78.5 degrees C, indicating that this lamellar-to-cubic phase transition is not accompanied by any significant volume change. For PLFE MLVs derived from cells grown at 65 degrees C, similar DSC and PPC profiles and thermal history responses were obtained. However, the lower growth temperature yields a higher DeltaV/V ( approximately 0.25%) and DeltaH (14 kJ/mol) value for the lamellar-to-lamellar phase transition measured at the same pH (2.1). A lower growth temperature also generates a less negative temperature dependence of alpha. The changes in DeltaV/V, DeltaH, and the temperature dependence of alpha can be attributed to the decrease in the number of cyclopentane rings in PLFE at the lower growth temperature. The relatively low DeltaV/V and small DeltaH involved in the phase transitions help to explain why PLFE liposomes are remarkably thermally stable and also echo the proposal that PLFE liposomes are generally rigid and tightly packed. These results help us to understand why, despite the occurrence of thermal-induced phase transitions, PLFE liposomes exhibit a remarkably low temperature sensitivity of proton permeation and dye leakage.  相似文献   

11.
Two types of hydrocarbon chain interdigitation in sphingomyelin bilayers   总被引:4,自引:0,他引:4  
Vibrational Raman spectroscopic experiments have been performed as a function of temperature on aqueous dispersions of synthetic DL-erythro-N-lignoceroylsphingosylphosphocholine [C(24):SPM], a racemic mixture of two highly asymmetric hydrocarbon chain length sphingomyelins. Raman spectral peak-height intensity ratios of vibrational transitions in the C-H stretching-mode region show that the C(24):SPM-H2O system undergoes two thermal phase transitions centered at 48.5 and 54.5 degrees C. Vibrational data for fully hydrated C(24):SPM are compared to those of highly asymmetric phosphatidylcholine dispersions. The Raman data are consistent with the plausible model that the lower temperature transition can be ascribed to the conversion of a mixed interdigitated gel state (gel II) to a partially interdigitated gel state (gel I) and that the higher temperature transition corresponds to a gel I----liquid-crystalline phase transition. The observation of a mixed interdigitated gel state (gel II) at temperatures below 48.5 degrees C implies that biological membranes may have lipid domains in which some of the lipid hydrocarbon chains penetrate completely across the entire hydrocarbon width of the lipid bilayer.  相似文献   

12.
The binding of polymyxin-B to lipid bilayer vesicles of synthetic phosphatidic acid was studied using fluorescence, ESR spectroscopy and electron microscopy. 1,6-Diphenylhexatriene (which exhibits polarized fluorescence) and pyrene decanoic acid (which forms excimers) were used as fluorescence probes to study the lipid phase transition. The polymyxin binds strongly to negatively charged lipid layers. As a result of lipid/polymyxin chain-chain interactions, the transition temperature of the lipid. This can be explained in terms of a slight expansion of the crystalline lipid lattice (Lindeman's rule). Upon addition of polymyxin to phosphatidic acid vesicles two rather sharp phase transitions (width deltaT = 5 degrees C) are observed. The upper transition (at Tu) is that of the pure lipid and the lower transition (at T1) concerns the lipid bound to the peptide. The sharpness of these transitions strongly indicates that the bilayer is characterized by a heterogeneous lateral distribution of free and bound lipid regions, one in the crystalline and the other in the fluid state. Such a domain structure was directly observed by electron microscopy (freeze etching technique). In (1 : 1) mixtures of dipalmitoyl phosphatidic acid and egg lecithin, polymyxin induces the formation of domains of charged lipid within the fluid regions of egg lecithin. With both fluorescence methods the fraction of lipid bound to polymyxin-B as a function of the peptide concentration was determined. S-shaped binding curves were obtained. The same type of binding curve is obtained for the interaction of Ca2+ with phosphatidic acid lamellae, while the binding of polylysine to such membranes is characterized by a linear or Langmuir type binding curve. The S-shaped binding curve can be explained in terms of a cooperative lipid-ligand (Ca2+, polymyxin) interaction. A model is proposed which explains the association of polymyxin within the membrane plane in terms of elastic forces caused by the elastic distortion of the (liquid crystalline) lipid layer by this highly asymmetric peptide.  相似文献   

13.
The behavior of phosphatidylethanolamine (PE) liposomes has been studied as a function of temperature, pH, ionic strength, lipid concentration, liposome size, and divalent cation concentration by differential scanning calorimetry (DSC), by light scattering, by assays measuring liposomal lipid mixing, contents mixing, and contents leakage, and by a new fluorometric assay for hexagonal (HII) transitions. Liposomes were either small or large unilamellar, or multilamellar. Stable (impermeable, nonaggregating) liposomes of egg PE (EPE) could be formed in isotonic saline (NaCl) only at high pH (greater than 8) or at lower pH in the presence of low ionic strength saline (less than 50 mOsm). Bilayer to hexagonal (HII) phase transitions and gel to liquid-crystalline transitions of centrifuged multilamellar liposomes were both detectable by DSC only at pH 7.4 and below. The HII transition temperature increased, and the transition enthalpy decreased, as the pH was raised above 7.4, and it disappeared above pH 8.3 where PE is sufficiently negatively charged. HII transitions could be detected at high pH following the addition of Ca2+ or Mg2+. No changes in light scattering and no lipid mixing, mixing of contents, or leakage of contents were noted for EPE liposomes under nonaggregating conditions (pH 9.2 and 100 mM Na+ or pH 7.4 and 5 mM Na+) as the temperature was raised through the HII transition region. However, when aggregation of the liposomes was induced by addition of Ca2+ or Mg2+, or by increasing [Na+], it produced sharp increases in light scattering and in leakage of contents and also changes in fluorescent probe behavior in the region of the HII transition temperature (TH). Lipid mixing and contents mixing were also observed below TH under conditions where liposomes were induced to aggregate, but without any appreciable leakage of contents. We conclude that HII transitions do not occur in liposomes under conditions where intermembrane contacts do not take place. Moreover, fusion of PE liposomes at a temperature below TH can be triggered by H+, Na+, Ca2+, or Mg2+ or by centrifugation under conditions that induce membrane contact. There was no evidence for the participation of HII transitions in these fusion events.  相似文献   

14.
We have studied the physical properties of aqueous dispersions of 1,2-sn- and 2,3-sn-didodecyl-beta-D-glucopyranosyl glycerols, as well as their diastereomeric mixture, using differential scanning calorimetry and low angle x-ray diffraction. Upon heating, both the chiral lipids and the diastereomeric mixture exhibit characteristically energetic L beta/L alpha phase transitions at 31.7-32.8 degrees C and two or three weakly energetic thermal events between 49 degrees C and 89 degrees C. In the diastereomeric mixture and the 1,2-sn glycerol derivative, these higher temperature endotherms correspond to the formation of, and interconversions between, several nonlamellar structures and have been assigned to L alpha/QIIa, QIIa/QIIb, and QIIb/HII phase transitions, respectively. The cubic phases QIIa and QIIb, whose cell lattice parameters are strongly temperature dependent, can be identified as belonging to space groups Ia3d and Pn3m/Pn3, respectively. In the equivalent 2,3-sn glucolipid, the QIIa phase is not observed and only two transitions are seen at 49 degrees C and 77 degrees C, which are identified as L alpha/QIIb and QIIb/HII phase transitions, respectively. These phase transitions temperatures are some 10 degrees C lower than those of the corresponding phase transitions observed in the diastereomeric mixture and the 1,2-sn glycerol derivative. On cooling, all three lipids exhibit a minor higher temperature exothermic event, which can be assigned to a HII/QIIb phase transition. An exothermic L alpha/L beta phase transition is observed at 30-31 degrees C. A shoulder is sometimes discernible on the high temperature side of the L alpha/L beta event, which may originate from a QIIb/L alpha phase transition prior to the freezing of the hydrocarbon chains. None of the lipids show evidence of a QIIa phase on cooling. No additional exothermic transitions are observed on further cooling to -3 degrees C. However, after nucleation at 0 degrees C followed by a short period of annealing at 22 degrees C, the 1,2-sn glucolipid forms an Lc phase that converts to an L alpha phase at 39.5 degrees C on heating. Neither the diastereomeric mixture nor the 2,3-sn glycerol derivative shows such behavior even after extended periods of annealing. Our results suggest that the differences in the phase behavior of these glycolipid isomers may not be attributable to headgroup size per se, but rather to differences in the stereochemistry of the lipid polar/apolar interfacial region, which consequently effects hydrogen-bonding, hydration, and the hydrophilic/hydrophobic balance.  相似文献   

15.
We investigated the effects of various concentrations of staphylococcal delta-lysin on the thermotropic phase behavior of large multilamellar dimyristoylphosphatidylcholine (DMPC) vesicles by differential scanning calorimetry (DSC), 31P nuclear magnetic resonance (NMR) and Fourier transform infrared (FTIR) spectroscopy, and X-ray diffraction. The DSC studies revealed that at all concentrations, the addition of delta-lysin progressively decreases the enthalpy of the pretransition of DMPC bilayers without significantly affecting its temperature or cooperativity. Similarly, the addition of smaller quantities of peptide has little effect on the temperature of the main phase transition of DMPC bilayers but does reduce the cooperativity and enthalpy of this transition somewhat. However, at higher peptide concentrations, a second phase transition with a slightly increased temperature and a markedly reduced cooperativity and enthalpy is also induced, and this latter phase transition resolves itself into two components at the highest peptide concentrations that are tested. Moreover, our 31P NMR spectroscopic studies reveal that at relatively low delta-lysin concentrations, essentially all of the phospholipid molecules produce spectra characteristic of the lamellar phase, whereas at the higher peptide concentrations, an increasing proportion exhibit an isotropic signal. Also, at the highest delta-lysin concentrations that are studied, the isotropic component of the 31P NMR spectrum also resolves itself into two components. At the highest peptide concentration that was tested, we are also able to effect a macroscopic separation of our sample into two fractions by centrifugation, a pellet containing relatively smaller amounts of delta-lysin and a supernatant containing larger amounts of peptide relative to the amount of lipid present. We are also able to show that the more cooperative phase transition detected calorimetrically, and the lamellar phase 31P NMR signal, arise from the pelleted material, while the less cooperative phase transition and the isotropic 31P NMR signal arise from the supernatant. In addition, we demonstrate by X-ray diffraction that the pelleted material corresponds to delta-lysin-containing large multilamellar vesicles and the supernatant to a mixture of delta-lysin-containing small unilamellar vesicles and discoidal particles. We also show by FTIR spectroscopy that delta-lysin exists predominantly in the alpha-helical conformation in aqueous solution or when interacting with DMPC, and that a large fraction of the peptide bonds undergo H-D exchange in D(2)O. However, upon interaction with DMPC, the fraction of exchangeable amide protons decreases. We also demonstrate by this technique that both of the phase transitions detected by DSC correspond to phospholipid hydrocarbon chain-melting phase transitions. Finally, we show by several techniques that the absolute concentrations of delta-lysin and the thermal history, as well as the lipid:peptide ratio, can affect the thermotropic phase behavior and morphology of peptide-lipid aggregates.  相似文献   

16.
The phase transition properties of aqueous suspensions of a series of nonhydrated (not heated above room temperature) and hydrated 1,2 diacylphosphatidylethanolamines (PE's) have been examined by high sensitivity differential scanning calorimetry at scan rates of 0.02-1.0 K min-1. At all scan rates nonhydrated PE's show a single asymmetric transition curve of excess heat capacity as a function of temperature. Multilamellar dispersions of hydrated PE's, however, exhibit transitions with fine structure, which can be fitted as the sum of three two-state component transitions, at scan rates of 0.02-0.1 K min-1, but give only a single asymmetric transition at 1.0 K min-1. At all scan rates the transition(s) of hydrated samples occur at lower temperatures than those of nonhydrated samples. One of the component transitions of hydrated PE's may be analogous to the pretransition that occurs in 1,2 diacylphosphatidylcholines.  相似文献   

17.
The bilayer to hexagonal phase transition of dioleoylphosphatidylethanolamine has been detected for the first time by differential scanning calorimetry. The observed transition is dependent on scan rate. This dependence can be explained by assuming that at rapid scan rates, the rate of conversion of bilayer to hexagonal phase is too slow at low temperatures for equilibration to take place. At higher temperatures the rate of interconversion becomes more rapid. The transition is observed to occur at 14°C using a scan rate of 0.74 K/min while it is centered at 8°C using a scan rate of 0.19 K/min. The enthalpy of the transition is 290 ± 40 cal/mol lipid and the transition is characterized by a ΔCp of −9 ± 1 mcal K−1 (g lipid)−1. The bilayer to hexagonal phase transition of dielaidoylphosphatidylethanolamine and of 1-palmitoy1-2-oleoylphosphatidylethanolamine occurs at 65.6°C and 71.4°C, respecitvely, with a corresponding transition enthalpy of 450 ± 20 and 400 ± 30 cal/mol lipid. The transitions of these phosphatidylethanolamines, occuring at higher temperatures, are independent of scan rate and show a higher degree of cooperativity than that of dioleoylphosphatidylethanolamine. Compared with the gel to liquid-crystalline transition of bilayer phospholipids the transition to hexagonal phase has a much lower enthalpy.  相似文献   

18.
The basic protein of myelin can spontaneously associate with the synthetic phospholipid N-palmitoyl-sphingosinephosphatidylcholine. The protein alters the phase transition properties of the lipid from a single transition at 41.5 degrees C to two overlapping transitions, one being slightly above and the other slightly below the transition temperature of the pure lipid. The effect was not seen upon the addition of poly(L-lysine) to this lipid nor does the myelin basic protein alter the phase transition properties of dimyristoylphosphatidylcholine. The results thus demonstrate that the myelin basic protein can interact with a major zwitterionic lipid component of myelin in addition to acidic phospholipids.  相似文献   

19.
Temperature induced denaturation of collagen in acidic solution   总被引:1,自引:0,他引:1  
Mu C  Li D  Lin W  Ding Y  Zhang G 《Biopolymers》2007,86(4):282-287
The denaturation of collagen solution in acetic acid has been investigated by using ultra-sensitive differential scanning calorimetry (US-DSC), circular dichroism (CD), and laser light scattering (LLS). US-DSC measurements reveal that the collagen exhibits a bimodal transition, i.e., there exists a shoulder transition before the major transition. Such a shoulder transition can recover from a cooling when the collagen is heated to a temperature below 35 degrees C. However, when the heating temperature is above 37 degrees C, both the shoulder and major transitions are irreversible. CD measurements demonstrate the content of triple helix slowly decreases with temperature at a temperature below 35 degrees C, but it drastically decreases at a higher temperature. Our experiments suggest that the shoulder transition and major transition arise from the defibrillation and denaturation of collagen, respectively. LLS measurements show the average hydrodynamic radius R(h), radius of gyration R(g)of the collagen gradually decrease before a sharp decrease at a higher temperature. Meanwhile, the ratio R(g)/R(h) gradually increases at a temperature below approximately 34 degrees C and drastically increases in the range 34-40 degrees C, further indicating the defibrillation of collagen before the denaturation.  相似文献   

20.
The basic protein of myelin can spontaneously associate with the synthetic phospholipid N-palmitoylsphingosinephosphatidylcholine. The protein alters the phase transition properties of the lipid from a single transition at 41.5°C to two overlapping transitions, one being slightly above and the other slightly below the transition temperature of the pure lipid. The effect was not seen upon the addition of poly(l-lysine) to this lipid nor does the myelin basic protein alter the phase transition properties of dimyristoylphosphatidylcholine. The results thus demonstrate that the myelin basic protein can interact with a major zwitterionic lipid component of myelin in addition to acidic phospholipids.  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