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 fluorescence emission properties of 1,6-diphenyl-1,3,5-hexatriene (DPH) in 1,2-dipalmitoyl-3-sn-phosphatidylcholine and 1,2-dimyristoyl-3-sn-phosphatidylcholine multilamellar vesicles have been measured by using multifrequency phase fluorometry. The fluorescence decay of DPH in the phospholipid vesicles has been analyzed by assuming either that the decay is made up of a discrete sum of exponential components or that the decay is made up of one or more continuous distributions of lifetime components. The fit of the decay curve using exponentials required at least two terms, and the reduced X2 was relatively large. The fit using a continuous distribution of lifetime values used two continuous components. Several symmetric distribution functions were used: uniform, Gaussian, and Lorentzian. The distribution function that best described the decay was the Lorentzian. The full width at half-maximum of the Lorentzian distribution was about 0.6 ns at temperatures below the phase transition temperature. At the phospholipid phase transition and at higher temperatures, the distribution became quite narrow, with a width of about 0.1 ns. It is proposed that the lifetime distribution is generated by a continuum of different environments of the DPH molecule characterized by different dielectric constants. Below the transition temperature in the gel phase, the dielectric constant gradient along the membrane normal determines the distribution of decay rates. Above the transition, in the liquid-crystalline phase, the translational and rotational mobility of the DPH molecule increases, and the DPH experiences an average environment during the excited-state lifetime. Consequently, the distribution becomes narrower.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The microviscosity of unilamellar vesicles of dimyristoyl-3-sn-phosphatidylcholine and that of phosphatidylcholine . apoprotein complexes was followed by fluorescence depolarization after labeling with 1,6-diphenyl-1,3,5-hexatriene. The transition temperature from gel-crystalline to liquid-crystalline phase in 24 degrees C for the dimyristoyl-phosphatidylcholine vesicles and is shifted to around 30 degrees C in the complexes between phosphatidylcholine and apoA-I, apoA-II, apoC-I, apoC-III proteins while the cooperativity of the transition is decreased. At temperatures below the transition of the phospholipid, the microviscosity of the complexes of phosphatidylcholine with apoA-I, apoA-II and apoC-I proteins is lower than that of the phosphatidylcholine, while the opposite effect is observed above 30 degrees C. The phosphatidylcholine . apoprotein complexes isolated on a Sepharose 6B column have a molecular weight around 100 000 and a phosphatidylcholine/apoprotein ratio of 2--2.6 (w/w). The microviscosity measurments at 35 degrees C performed after elution of the column enable the complex to be detected. The size and microviscosity of the apoprotein . phosphatidylcholine complex is compatible with a model where the vesicular structure has disappeared and the amino acid side chains present hydrophobic interaction with the phosphatidylcholine acyl chains.  相似文献   

3.
Prothrombin denaturation was examined in the presence of Na2EDTA, 5mM CaCl2, and CaCl2 plus membranes containing 1-palmitoyl-2-oleoyl-3-sn-phosphatidylcholine (POPC) in combination with either bovine brain phosphatidylserine (PS) or 1,2-dioleoyl-phosphatidylglycerol (DOPG). Heating denaturation of prothrombin produced thermograms showing two peaks, a minor one at approximately 59 degrees C previously reported to correspond to denaturation of the fragment 1 region (Ploplis, V. A., D. K. Strickland, and F. J. Castellino 1981. Biochemistry. 20:15-21), and a main one at approximately 57-58 degrees C, reportedly due to denaturation of the rest of the molecule (prethrombin 1). The main peak was insensitive to the presence of 5mM Ca2+ whereas the minor peak was shifted to higher temperature (Tm approximately 65 degrees C) by Ca2+. Sufficient concentrations of POPC/bovPS (75/25) large unilamellar vesicles to guarantee binding of 95% of prothrombin resulted in an enthalpy loss in the main endotherm and a comparable enthalpy gain in the minor endotherm accompanying an upward shift in peak temperature (Tm approximately 73 degrees C). Peak deconvolution analysis on the prothrombin denaturation profile and comparison with isolated prothrombin fragment 1 denaturation endotherms suggested that the change caused by POPC/PS vesicles reflected a shift of a portion of the enthalpy of the prethrombin 1 domain to higher temperature (Tm approximately 77 degrees C). The enthalpy associated with this high-temperature endotherm increased in proportion to the surface concentration of PS. By contrast, POPC/DOPG (50/50) membranes shifted the prethrombin 1 peak by 4 degrees C to a lower temperature and the fragment 1 peak by 5 degrees C to a higher temperature. The data lead to a hypothesis that the fragment 1 and prethrombin 1 domains of prothrombin do not denature quite independently and that binding of prothrombin to acidic-lipid membranes disrupts the interaction between these domains. It is further hypothesized that PS containing membranes exert the additional specific effect of decoupling the denaturation of two subdomains of the prethrombin 1 domain of prothrombin.  相似文献   

4.
Fourier transform infrared spectroscopy was used to study the metastability of 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) at temperatures near 0 degrees C. It was found that when DPPC is incubated at 2 degrees C for three days the two-dimensional acyl chain packing changes from one resulting in spectra typical of an orthorhombic subcell to one resembling that found in triclinically packed acyl systems. This transition proceeds in two stages. The first step, requiring less than one day, approximates first-order kinetics; the second stage proceeds with second- or higher-order kinetics. Comparison of spectra recorded at -36 degrees C with and without prior incubation at 2 degrees C shows that there are two stable low temperature forms of DPPC; that is, DPPC is metastable only within a narrow temperature range. A study of the thermotropic behavior in the range 0-45 degrees C shows that the subtransition near 15 degrees C is a transition from the alternate form to one with orthorhombic characteristics. Spectral changes at the pretransition and the main phase transition demonstrate that there are differences in behavior that are related to the thermal history of the sample.  相似文献   

5.
Unilamellar vesicles of varying and reasonably uniform size were prepared from 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) by the extrusion procedure and sonication. Quasi-elastic light scattering was used to show that different vesicle preparations had mean (Z-averaged) diameters of 1340, 900, 770, 630, and 358 A (sonicated). Bilayer-phase behavior as detected by differential scanning calorimetry was consistent with the existence of essentially uniform vesicle populations of different sizes. The response of these different vesicles to treatment with poly(ethylene glycol) (PEG) was monitored using fluorescence assays for lipid transfer, contents leakage, and contents mixing, as well as quasi-elastic light scattering. No fusion, as judged by vesicle contents mixing and change in vesicle size, was detected for vesicles of diameter greater than 770 A. The diameters of smaller vesicles increased dramatically when treated with high concentrations of PEG, although mixing of their contents could not be detected both because of their small trapped volumes and because of the extensive leakage induced in small vesicles by high concentrations of PEG. Lipid transfer was detected between vesicles of all sizes. We conclude the high bilayer curvature does encourage fusion of closely juxtaposed membrane bilayers but that highly curved vesicles appear also to rupture and form larger structures when diluted from high PEG concentration, a process that can be confused with fusion. Despite the failure of PEG to induce fusion of large, uncurved vesicles composed of a single phosphatidylcholine, these vesicles can be induced to fuse when they contain small amounts of certain amphiphathic compounds thought to play a role in cellular fusion processes. Thus, vesicles which contained 0.5 mol % L-alpha-lysopalmitoylphosphatidylcholine, 5 mol % platelet activating factor, or 0.5 mol % palmitic acid fused in the presence of 30%, 25%, and 20% (w/w) PEG, respectively. However, vesicles containing 1,2-dipalmitoyl-sn-glycerol, 1,2-dioleoyl-sn-glycerol, 1-oleoyl-2-acetyl-sn-glycerol, or monooleoyl-rac-glycerol at surface concentrations up to 5 mol % did not fuse in the presence or absence of PEG. There was no correlation between the abilities of these amphipaths to induce phase separation or nonlamellar phases and their abilities to support fusion of pure DPPC unilamellar vesicles in the presence of high concentrations of PEG. The results are discussed in terms of the type of disrupted lipid packing that could be expected to favor PEG-mediated fusion.  相似文献   

6.
Properties of large unilamellar vesicles (LUV), composed of phosphatidylcholine and prepared by reverse-phase evaporation and subsequent extrusion through Unipore polycarbonate membranes, have been investigated and compared with those of small unilamellar vesicles (SUV) and of multilamellar vesicles (MLV). The unilamellar nature of the LUV is shown by 1H-NMR using Pr3+ as a shift reagent. The gel to liquid-crystalline phase transition of LUV composed of dipalmitoylphosphatidylcholine (DPPC) monitored by differential scanning calorimetry, fluorescence polarization of diphenylhexatriene and 90 degrees light scattering, occurs at a slight lower temperature (40.8 degrees C) than that of MLV (42 degrees C) and is broadened by about 50%. The phase transition of SUV is shifted to considerably lower temperatures (mid-point, 38 degrees C) and extends over a wide temperature range. In LUV a well-defined pretransition is not observed. The permeability of LUV (DPPC) monitored by leakage of carboxyfluorescein, increases sharply at the phase transition temperature, and the extent of release is greater than that from MLV. Leakage from SUV occurs in a wide temperature range. Freeze-fracture electron microscopy of LUV (DPPC) reveals vesicles of 0.1-0.2 micron diameter with mostly smooth fracture faces. At temperatures below the phase transition, the larger vesicles in the population have angled faces, as do extruded MLV. A banded pattern, seen in MLV at temperatures between the pretransition and the main transition, is not observed in the smaller LUV, although the larger vesicles reveal a dimpled appearance.  相似文献   

7.
《Biophysical journal》2020,118(10):2411-2425
In this study, we investigated the conversion of lipid bicelles into vesicles in the case of a system composed of the phospholipid 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and the saponin glycyrrhizin in the presence of sucrose. Glycyrrhizin is a biosurfactant present in the licorice root and possesses a triterpenic hydrophobic backbone and a hydrophilic headgroup built from two sugar molecules. The aim of this study is to determine the initial bicelle size at temperatures below the lipid’s main phase transition temperature Tm and, based on these results, characteristics of the temperature-induced bicelle-to-vesicle transition. Moreover, the influence of the heating rate on this transition is followed. The general picture concluded from photon correlation spectroscopy and small angle X-ray scattering was confirmed by additional imaging with cryogenic transmission electron microscopy. Small angle X-ray scattering was especially used to determine size parameters of the existing structures. To enhance the contrast for X-rays, a buffer containing 25 wt% sucrose was used. It was found that larger vesicles were formed from smaller precursor particles and that monodisperse precursors are required for formation of very monodisperse vesicles upon temperature increase. At high glycyrrhizin contents and above a critical heating rate of ∼5°C min−1, the polydispersity of these vesicles is decoupled from both parameters, glycyrrhizin content and heating rate. However, the vesicle size stays tunable by the glycyrrhizin content and increases upon increasing the glycyrrhizin concentration. Therefore, vesicles of defined size and with a rather low polydispersity of ∼12–14% can be formed.  相似文献   

8.
A W Scotto  D Zakim 《Biochemistry》1986,25(7):1555-1561
The presence of cholesterol in small unilamellar vesicles (ULV) of dimyristoylphosphatidylcholine (DMPC) catalyzes fusion of the vesicles at temperatures below the upper limit for the gel to liquid-crystalline phase transition of the DMPC. The extent to which ULV grow depends on the concentration of cholesterol in the vesicles and on temperature. Maximum growth occurs at 21 degrees C. It decreases as the temperature is lowered below 21 degrees C. Growth does not occur at temperatures above the phase transition. In addition, the presence of cholesterol in ULV of DMPC catalyzes the insertion of integral membrane proteins into the vesicles. Thus, bacteriorhodopsin from Halobacterium halobrium, UDPglucuronosyltransferase (EC 2.4.1.17) from pig liver microsomes, and cytochrome oxidase from beef heart mitochondria formed stable lipid-protein complexes spontaneously when added to ULV containing cholesterol at temperatures under which these vesicles would fuse. Incorporation of these proteins into the ULV of DMPC did not occur in the absence of cholesterol or in the presence of cholesterol when the temperature of the system was above that for the phase transition. It appears that cholesterol lowers the energy barrier for fusion of ULV of DMPC and for insertion of integral membrane proteins into these bilayers. Studies with bacteriorhodopsin suggest that the energy barrier for insertion of proteins into ULV containing cholesterol is smaller than the energy barrier for fusion of the ULV with each other.  相似文献   

9.
The interaction of diacylglycerols, primarily 1,2-dilauroyl-sn-glycerol (1,2-DLG), with egg phosphatidylcholine (PC) bilayers was studied by NMR spectroscopy and other physical techniques. In the low proportions used (less than or equal to 20 mol % with respect to total lipid), 1,2-DLG formed bilayers with PC with no hexagonal phase separation, as assessed by light, polarizing and electron microscopy, and 31P and 13C NMR spectroscopy. The 13C-carbonyl chemical shift of 90% [13C]carbonyl 1,2-DLG was monitored in small unilamellar vesicles as a function of relative DLG content (1.5-20%) and temperature (10-55 degrees C). The chemically inequivalent sn-1 and sn-2 carbonyls gave a single, narrow resonance in vesicles, in contrast to neat 1,2-DLG and 1,2-DLG in organic solvents, whose spectra showed two well-separated carbonyl resonances. The chemical shift of 1,2-DLG in PC shows that the carbonyl groups are proximal to the aqueous interface, necessitating orientation of the DLG molecule along the normal to the bilayer. Both carbonyl groups are H-bonded to H2O, but the secondary ester (sn-2) carbonyl is relatively more hydrated than the primary ester (sn-1) carbonyl. The 13C-carbonyl chemical shift data further suggest that the interfacial conformation resembles that of crystalline and liquid crystalline lamellar 1,2-dilauroyl-sn-glycero-3-phosphatidylethanolamine and certain PCs, in which the glycerol backbone is perpendicular to the bilayer plane. This conformation is different from that of crystalline 1,2-dilauroyl-sn-glycerol, in which the glycerol backbone is parallel to the bilayer plane. Between 1.5 and 8% DLG in vesicles, the chemical shift of the 1,2-DLG carbonyl at a given temperature was constant. However, above 8% DLG the chemical shift at each temperature increased with increasing DLG concentration, suggesting increased hydration at higher DLG content. At low temperatures 13C NMR spectra of vesicles with the highest proportions of 1,2-DLG studied (15 and 20%) showed two DLG carbonyl resonances, which most likely represent 1,2-DLG on outer and inner leaflets of the vesicle bilayer. The two peaks collapsed into a single resonance by 38 degrees C, at which temperature the two environments equilibrate with a rate constant of approximately 60 s-1 (t1/2 approximately 10 ms). Thus, transbilayer movement of DLG is extremely fast compared with phospholipids. In vesicles the 1,3-isomer of DLG exhibited a narrow carbonyl peak slightly downfield from that of 1,2-DLG. Acyl chain migration from 1,2-DLG to 1,3-DLG was monitored directly in the vesicle by time-dependent NMR measurements.  相似文献   

10.
The permeability of egg yolk lecithin (EYL) vesicles to Pr3+ has been measured by 31P nuclear magnetic resonance (nmr) spectroscopy. Measurable Pr3+ leakage into the internal aqueous compartment of EYL vesicles at ambient (21 degrees C) temperature required the presence of small (7--10 mol%) amounts of dicetyl phosphate (DCP). The permeability of DCP-containing vesicles is decreased by incorporation of sterol (cholesterol greater than ergosterol approximately 5.6-dihydroergosterol greater than zymosterol) into the lipid bilayer. Addition of the polyene macrolide antibiotic, nystatin, to DCP-containing EYL vesicles with and without sterol resulted in increased Pr3+ permeability at the three temperatures studied (21--37.5 degrees C). Permeability changes observed upon addition of nystatin to sterol-impregnated, DCP-containing vesicles varied with sterol structure: ergosterol approximately 5,6-dihydroergosterol greater than cholesterol approximately zymosterol. These results are compared with other polyene macrolide induced permeability changes on model and natural membrane systems. Permeability changes induced by nystatin in sterol-free EYL vesicles were generally greater than for comparable sterol-containing vesicles. This is attributed to a nonspecific interaction of the antibiotic with the latter vesicles.  相似文献   

11.
Glycophorin from the human erythrocyte membrane has been isolated in pure form and reconstituted into large unilamellar vesicles comprised of binary mixtures of 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) and chain perdeuterated 1,2-dimyristoyl-3-sn-phosphatidylcholine (DMPC-d54). The effect of temperature and protein on lipid structure and mixing was monitored by using Fourier transform infrared spectroscopy; deuteration of one of the components of the mixture permits observation of the protein interaction with each lipid species. The melting curves were analyzed by assuming that each lipid chain can exist in one of two physical states (i.e., gel or liquid crystalline), characterized by a temperature-dependent Lorentzian distribution for the line shape of the C-H or C-D stretching vibrations. The fraction of each lipid component melted at temperatures within the two-phase region of the phase diagram was calculated and approximate phase diagrams were constructed. Addition of protein lowers the liquidus line of the phase diagram while leaving the solidus line essentially unchanged. No lipid phase separation is observed. The effect of protein is more pronounced on the DPPC component than on the DMPC-d54. The former is significantly more disordered and/or fluidized at all lipid mole fractions in the ternary system than in the binary phospholipid mixture.  相似文献   

12.
Solid-state 31P- and 13C-NMR spectra were recorded in melittin-lecithin vesicles composed of 1,2-dilauroyl-sn-glycero-3-phosphocholine (DLPC) or 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC). Highly ordered magnetic alignments were achieved with the membrane surface parallel to the magnetic field above the gel-to-liquid crystalline phase transition temperature (Tc). Using these magnetically oriented vesicle systems, dynamic structures of melittin bound to the vesicles were investigated by analyzing the 13C anisotropic and isotropic chemical shifts of selectively 13C-labeled carbonyl carbons of melittin under the static and magic-angle spinning conditions. These results indicate that melittin molecules adopt an alpha-helical structure and laterally diffuse to rotate rapidly around the membrane normal with tilt angles of the N-terminal helix being -33 degrees and -36 degrees and those of the C-terminal helix being 21 degrees and 25 degrees for DLPC and DPPC vesicles, respectively. The rotational-echo double-resonance method was used to measure the interatomic distance between [1-13C]Val8 and [15N]Leu13 to further identify the bending alpha-helical structure of melittin to possess the interhelical angles of 126 degrees and 119 degrees in DLPC and DPPC membranes, respectively. These analyses further lead to the conclusion that the alpha-helices of melittin molecules penetrate the hydrophobic cores of the bilayers incompletely as a pseudo-trans-membrane structure and induce fusion and disruption of vesicles.  相似文献   

13.
The effect of temperature on the membranes of Ant-300, a psychrophilic marine bacterium, was studied by measuring alanine uptake by isolated membrane vesicles. Uptake was observed from 0 to 35 degrees C. The maximum initial rate of uptake occurred at 25 degrees C although more alanine was ultimately taken up at temperatures from 10 to 20 degrees C. An ARRHENIUS plot of these data shows a single infection point at 7.8 degrees C. Within 10 min, over 50% of the alpha-aminoisobutyric acid taken up by whole cells at 5 degrees C was lost after a temperature shift to 25 degrees C. Vesicles preloaded with alanine at 5 degrees C did not become leaky when shifted to 25 degrees C. In addition, exposure of the vesicles to 25 degrees C for 30 min did not affect subsequent alanine uptake at 5 degrees C. The data obtained suggest that the loss of the uptake and permeability control functions of membranes from psychrophilic bacteria at elevated temperatures is not due to degeneration of the membrane itself, but rather to a control or regulatory mechanism associated with whole cells.  相似文献   

14.
We are aiming to improve the encapsulation efficiency of proteins in a size-regulated phospholipid vesicle using an extrusion method. Mixed lipids (1,2-dipalmitoyl-sn-glycero-3-phosphatidylcholine (DPPC), cholesterol, 1,5-dipalmitoyl-l-glutamate-N-succinic acid (DPEA), and 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[monomethoxy poly(ethylene glycol) (5,000)] (PEG-DSPE) at a molar ratio of 5, 5, 1, and 0.033 were hydrated with a NaOH solution (7.6 mM) to obtain a polydispersed multilamellar vesicle dispersion (50 nm to 30 microm diameter). The polydispersed vesicles were converted to smaller vesicles having an average diameter of ca. 500 nm with a relatively narrow size distribution by freeze-thawing at a lipid concentration of 2 g dL(-)(1) and cooling rate of -140 degrees C min(-1). The lyophilized powder of the freeze-thawed vesicles was rehydrated into a concentrated protein solution (carbonyl hemoglobin solution, 40 g dL(-1)) and retained the size and size distribution of the original vesicles. The resulting vesicle dispersion smoothly permeated through the membrane filters during extrusion. The average permeation rate of the freeze-thawed vesicles was ca. 30 times faster than that of simple hydrated vesicles. During the extrusion process, proteins were encapsulated into the reconstructed vesicles with a diameter of 250 +/- 20 nm.  相似文献   

15.
A Kintanar  A C Kunwar  E Oldfield 《Biochemistry》1986,25(21):6517-6524
We have investigated the deuterium (2H) nuclear magnetic resonance (NMR) spectra of two 2H-labeled fluorescence probes (trans,trans,trans-1,6-diphenylhexa-1,3,5-trienes, DPHs) incorporated into model lipid bilayer membrane systems at various temperatures. The membranes consisted of multilamellar bilayers of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) containing varying concentrations of cholesterol. The conventional one-order parameter approach often used in the analysis of the NMR data of lipid membranes does not explain the observed temperature variations of the spectral features. Consistent with the molecular symmetry, the results have thus been analyzed in terms of an ordering matrix with more than one independent element. The molecular order parameter (SNMR), the order along the long molecular axis, in the pure lipid system varies from 0.49 to 0.26 as the temperature is increased from 25 to 57 degrees C. These values are somewhat larger than the order parameters obtained from fluorescence depolarization (SFLU) on sonicated DMPC vesicles. Such discrepancies probably arise from the looser packing of the sonicated vesicles. Addition of cholesterol to the model membranes causes the order parameter of the probe molecules to increase. At 35 degrees C, SNMR increases from 0.38 (with no cholesterol) to 0.92 (in the presence of 50 mol % cholesterol). These values are about 10% larger than those obtained from fluorescence depolarization studies on sonicated vesicles. The SNMR for DPH are somewhat larger than those obtained in earlier NMR studies of 2H-labeled cholesterol. However, they compare well with those obtained for 2H-labeled DMPC.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Sonicated liposomes of dipalmitoyl phosphatidylcholine show sharp turbidity changes on heating at two distinct temperatures. A decrease in turbidity at the lower temperature (approx. 37°C) is thought to be associated with the phase transition of small vesicles and a decrease at about 44°C with larger vesicles or multilayer. An increase of turbidity between 38 and 43°C is attributed to the fusion of small vesicles. The turbidity changes were studied under various modes of vesicle preparation to confirm the interpretation of the turbidity data. Alternate interpretations are discussed.  相似文献   

17.
Acclimation of winter oilseed plants in the cold (i.e. at temperatures >0 degrees C) followed by short exposure to sub-lethal freezing temperatures resulted in pronounced ultrastructural changes of leaf epidermal and mesophyll cells. The following major changes were observed upon acclimation at 2 degrees C: increased thickness of cell walls; numerous invaginations of plasma membranes; the appearance of many large vesicles localized in the cytoplasm in close proximity to the central vacuole; the occurrence of abundant populations of microvesicles associated with the endoplasmic reticulum (ER) cisternae or located in the vicinity of dictyosomes; and the occurrence of paramural bodies and myelin-like structures. In addition, large phenolic deposits were observed in the vicinity of the plasma membrane and membrane-bound organelles such as chloroplasts, large vesicles or cytoplasm/tonoplast interfaces. Transient freezing (-5 degrees C for 18 h) of the cold-acclimated leaves led to reversible disorganization of the cytoplasm and to pronounced structural changes of the cellular organelles. Chloroplasts were swollen, with the stroma occupying one half of their volume and the thylakoid system being displaced to the other half. Large phenolic aggregates disappeared but distinct layers of phenolic deposits were associated with mitochondrial membranes and with chloroplast envelopes. In frost-thawed cells recovered at 2 degrees C for 24 h, dictyosomes and dictyosome- or ER-derived small vesicles reappeared in the ribosome-rich cytoplasm. Aberrations in the structure of chloroplasts and mitochondria were less pronounced. Few phenolic deposits were seen as small grains associated with chloroplast envelopes and vesicle membranes. These observations demonstrate that plants undergo different changes in cell ultrastructure depending on whether they are subjected to chilling or freezing temperatures. Results are discussed in relation to membrane recycling and the possible role of phenolics during the first and second stages of plant acclimation at low temperature.  相似文献   

18.
Membrane fusion is an obligatory step in many vital cellular processes. The well-established enrichment of bilayer-destabilizing lipids in membranes of poikilotherms subjected to growth at low temperatures leads to the prediction that such membranes will possess a greater propensity to undergo fusion. This hypothesis was explicitly tested in the present study by determining the kinetics of fusion between small unilamellar vesicles (SUVs) prepared from endoplasmic reticulum (ER) membranes of thermally-acclimated (to 5 and 20 degrees C) rainbow trout (Oncorhynchus mykiss) liver and bovine brain phosphatidylserine (BBPS). At temperatures above 10 degrees C, ER vesicles from 5 degrees C-acclimated trout, fused more rapidly and to a greater extent with BBPS vesicles (by average factors of 1.25- and 1.45-fold, respectively) than ER vesicles of 20 degrees C-acclimated trout. At temperatures >35 degrees C, apparent fusion rates declined while the extent of fusion increased in both acclimation groups. Fusion kinetics were found to be well correlated with and limited by the physical properties and phase state of the BBPS vesicles. These results indicate that dynamic attributes of biological membranes, such as the propensity to undergo fusion, are of potential regulatory significance and are partially conserved when growth or environmental temperature changes.  相似文献   

19.
A change in ambient temperature caused marked alterations, over a 24h period, in the proportions of the unsaturated C18 fatty acids in 3-sn-phosphatidylcholine and 1,2-diacyglycerols during the development of soya-bean and linseed cotyledons. The molar proportion of oleate increased when the temperature was increased whereas that of linoleate or linolenate, depending on the species, increased when the temperature was lowered. Concomitant changes in the composition of 3-sn-phosphatidylethanolamine and triacylglycerols were small.  相似文献   

20.
Large unilamellar vesicles of dipalmitoylphosphatidylcholine (DPPC) were prepared by sonication and were fractionated by gel filtration on Sepharose Cl-2B in the size range from 180- to 380-A Stokes radii. Negatively stained electron micrographs of these preparations indicated the presence of unilamellar, spheroidal structures of the expected size. Fluorescence polarization of diphenylhexatriene, dissolved in the vesicles, revealed progressively broader phase transitions, shifted to lower temperatures for vesicles of decreasing sizes. The fractionated unilamellar vesicles and multilamellar vesicles of DPPC were reacted with human apolipoprotein A-I at 41 degrees C for periods from 1 to 120 h. The reaction mixtures were then passed through a Bio-Gel A-5m column to separate unreacted lipid vesicles and protein from micellar complexes of DPPC with apolipoprotein A-I. Smaller vesicles were much more reactive than larger vesicles or multilamellar vesicles with the apolipoprotein. This difference in reactivity was explained by the increasing bilayer curvature of smaller vesicles which changes the packing of DPPC molecules in the bilayer and facilitates its penetration by the apolipoprotein.  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