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.
We have investigated the effect of two monosaccharides, glucose and fructose, and two disaccharides, sucrose and trehalose, on the thermotropic phase transition of unilamellar extruded vesicles of DPPC. All the sugars investigated raise the main transition temperature (Tm) of some fraction of the lipid, but there are differences between the effect of glucose and the other three sugars. At low concentrations of glucose, Tm is lowered. At high concentrations of glucose there are two transitions, one with a low Tm and one with a high Tm. The data suggest that at low concentrations, all of the glucose present may bind to the bilayer and increase headgroup spacing by physical intercalation or increased hydration. The appearance of a Tm above that of pure hydrated DPPC suggests the possibility of the dehydration of some other population of phospholipid molecules. The other three sugars increase Tm, but at high concentrations of trehalose, sucrose, and fructose a second peak occurs at a low Tm. The other sugars appear to dehydrate the bilayer at low concentrations, but may show some binding or increased hydration of some portion of the lipid at very high concentrations. The sugar effects on unilamellar vesicles are strikingly different from the effects of these sugars on multilamellar vesicles.  相似文献   

2.
Liposomes are commonly used as models for chilling and freezing damage, with leakage of water-soluble contents from the aqueous interior as the most frequently used measurement of damage. In order to achieve an understanding of the mechanism of the leakage, we have conducted a study of the factors that influence the leakage from liposomes during phase transitions. While such investigations have appeared sporadically in the literature, a detailed study has not been undertaken previously, despite the fact that liposomes are widely used as models for stress injury. Thus, we suggest that these findings will be of general interest in the cryobiology community. We now report that the following variables affected leakage from liposomes during chilling: (i) increasing the rate of cooling and warming resulted in decreased leakage; (ii) maximal leakage occurred at the measured phase transition temperature; (iii) addition of defect-forming additives such as a second phospholipid or a surfactant increased leakage from the liposomes during the phase transition but not above or below that temperature; (iv) small unilamellar vesicles leaked much more rapidly than large unilamellar vesicles; and (v) increasing the pH of the external buffer decreased leakage of carboxyfluorescein, an effect that is probably particular to ionizable solutes.  相似文献   

3.
Pullulan, a relationship between molecular weight and fine structure   总被引:9,自引:0,他引:9  
Catley BJ 《FEBS letters》1970,10(3):190-193
  相似文献   

4.
E H Ball  T Kovala 《Biochemistry》1988,27(16):6093-6098
Caldesmon is a widely distributed contractile protein that occurs in both a high molecular weight [120-150-kilodalton (kDa)] and a low molecular weight (71-80-kDa) form, depending on the tissue. The structural relationship between these two forms was examined by mapping techniques. Partial cyanogen bromide cleavage in conjunction with sodium dodecyl sulfate gel electrophoresis was used to construct a map of the cleavage points and determine the relative position of the fragments in a high molecular weight caldesmon from chicken gizzard (caldesmon125). By use of this map, markers for different regions of the protein were obtained: Antibodies directed toward certain areas were prepared by affinity purification, and specific 125I-labeled tryptic peptides were found to originate from terminal cyanogen bromide fragments. Mapping of a lower molecular weight form of caldesmon (caldesmon72 from chicken liver) revealed the presence of sequences located in both ends of caldesmon125. A terminal 38-kDa fragment of both proteins was apparently identical on the basis of arrangement of cleavage sites, antibody reactivity, and iodopeptide mapping. Fragments from the other end of both proteins exhibited an identical pattern of peptides. These results show that it is sequences located in the central area of caldesmon125 which are missing in caldesmon72, indicating that the smaller molecule is not simply a proteolytic product of the larger. The two forms of caldesmon may be derived from separate genes or by alternative splicing from a single gene.  相似文献   

5.
ABSTRACT: BACKGROUND: Body weight is at least partly controlled by the choices made by a human in response to external stimuli. Changes in body weight are mainly caused by energy intake. By analyzing the mechanisms involved in food intake, we considered that molecular diffusion plays an important role in body weight changes. We propose a model based on Fick's second law of diffusion to simulate the relationship between energy intake and body weight. RESULTS: This model was applied to food intake and body weight data recorded in humans; the model showed a good fit to the experimental data. This model was also effective in predicting future body weight. CONCLUSIONS: In conclusion, this model based on molecular diffusion provides a new insight into the body weight mechanisms.  相似文献   

6.
  • 1.1. The following equation based on 230 conductance values for 192 species of mammals of body weights ranging from 3.5 to 150,000 g describes the relationship of conductance below thermal neutrality to body weight in mammals: C = 0.760 W−0.426, where C has units of mlO2/g·h·°C and W is body weight in g.
  • 2.2. Bats, order Chiroptera, have conductance values higher than predicted from body weight; conductance is predicted by the equation : C = 1.54 W−0.54.
  • 3.3. Heteromyid and cricetid rodents have conductance values below predicted and the following equations predict conductance in these two families. C = 0.62 W−0.44 and C = 1.03 W−0.54, respectively.
  相似文献   

7.
The relationship between the molecular weight and the number of subunits in oligomeric globular proteins consisting of identical subunits has been analyzed. It has been shown that the molecular weights of the subunits are distributed about a mean value of 48,000 and consequently that the molecular weights of the native oligomeric proteins are distributed in clearly distinguishable molecular weight regions. This observation allows the probability of a particular oligomeric structure to be predicted from a measurement of the oligomer molecular weight alone, which is useful in a number of types of study of protein structure, particularly comparative studies. Calculations have been performed which suggest that there is no thermodynamic limitation, in terms of the subunit interactions themselves, to the size of an oligomeric protein with a given number of subunits. Rather, an individual polypeptide chain itself has inherent size limitations, which consequently limits the molecular weight of the corresponding oligomer.  相似文献   

8.
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   

9.
10.
Thermotropic transitions of dihexadecylphosphatidylcholine (DHPC) dispersions in hydrogen oxide (1H2O) and deuterium oxide (2H2O) were investigated by differential scanning calorimetry (DSC). In DHPC dispersions, transition temperature between interdigitated gel phase (L beta I) and ripple phase (P beta') is lower in 2H2O than in 1H2O, and transition between the ripple phase (P beta') and fluid phase (L alpha) in 2H2O occurs at a temperature slightly higher than in 1H2O. In dipalmitoylphosphatidylcholine (DPPC) dispersions, on the other hand, transition temperature between lamellar gel phase (L beta') and ripple phase is higher in 2H2O than in 1H2O. These results suggest that the interdigitated gel phase is more stable in 1H2O than in 2H2O. To account for the shift of transition temperature by the water substitution, difference of interfacial energies between these aqueous environments is discussed.  相似文献   

11.
The interaction of La3+ with phosphatidylserine vesicles is elucidated by binding studies, differential scanning calorimetry, X-ray diffraction, freeze fracture electron microscopy, and release of vesicle contents. La3+ effectively competes with Ca2+ for phosphatidylserine binding sites. The saturation level is close to a La/lipid ratio of 1:3. A concentration of 0.1 mM of La3+ is sufficient to induce fusion between sonicated vesicles.  相似文献   

12.
In this review, we summarize the results of recent studies on the main phase transition behavior of phospholipid bilayers using the combined approaches of molecular mechanics simulations and high-resolution differential scanning calorimetry. Following a brief overview of the phase transition phenomenon exhibited by the lipid bilayer, we begin with the review by showing how several structural parameters underlying various phospholipids including phosphatidylcholine, phosphatidylethanolamine, and phosphatidylglycerol are defined and determined. Specifically, these structural parameters are obtained with saturated lipids packed in the gel-state bilayer using computer-based molecular mechanics calculations. Then we proceed to present the calorimetric data obtained with the lipid bilayer composed of saturated phospholipids as it undergoes the gel-to-liquid-crystalline phase transition in excess water. The general equations that can correlate the gel-to-liquid-crystalline phase transition temperature (Tm) of the lipid bilayer with the structural parameters of the lipid molecule constituting the lipid bilayer are subsequently presented. From these equations, two tables of predicated Tm values for well over 400 molecular species of saturated phosphatidylcholine and saturated phosphatidylethanolamine are generated. We further review the structure and chain-melting behavior of a large number of sn-1 saturated/sn-2 unsaturated phospholipids. Two Tm-diagrams are shown, from which the effects of the number and the position of one to five cis carbon–carbon double bonds on Tm can be viewed simultaneously. Finally, in the last part of this review, simple molecular models that have been invoked to interpret the characteristic Tm trends exhibited by lipid bilayers composed of unsaturated lipids with different numbers and positions of cis carbon–carbon double bonds as seen in the Tm-diagram are presented.  相似文献   

13.
H Ellens  J Bentz  F C Szoka 《Biochemistry》1986,25(14):4141-4147
The initial kinetics of fusion and leakage of liposomes composed of N-methylated dioleoylphosphatidylethanolamine (DOPE-Me) have been correlated with the phase behavior of this lipid. Gagné et al. [Gagné, J., Stamatatos, L., Diacovo, T., Hui, S. W., Yeagle, P., & Silvius, J. (1985) Biochemistry 24, 4400-4408] have shown that this lipid is lamellar (L alpha) below 20 degrees C, is hexagonal (HII) above 70 degrees C, and shows isotropic 31P NMR resonances at intermediate temperatures. This isotropic state is also characterized by complex morphological structures. We have prepared DOPE-Me liposomes at pH 9.5 and monitored the temperature dependence of the mixing of aqueous contents, leakage, and changes in light scattering upon reduction of the pH to 4.5. At and below 20 degrees C, where the lipid is in the L alpha phase, there is very little aggregation or destabilization of the liposomes. Between 30 and 60 degrees C, i.e., where the lipid is in the isotropic state, the initial rates of liposome fusion (mixing of aqueous contents) and leakage increase. At temperatures approaching that where the hexagonal HII phase transition occurs, the initial rates and extents of fusion decrease, whereas leakage is enhanced. Similar results were found for dioleoylphosphatidylethanolamine/dioleoylphosphatidylcholine (2:1) liposomes. These results clearly establish a common mechanism between the appearance of the isotropic state (between the L alpha and HII phases) and the promotion of liposome fusion. We propose a simple model to explain both the observed behavior of phosphatidylethanolamine-containing membranes with respect to liposome fusion and/or lysis and the beginning of the L alpha-HII phase transition.  相似文献   

14.
In this review, we summarize the results of recent studies on the main phase transition behavior of phospholipid bilayers using the combined approaches of molecular mechanics simulations and high-resolution differential scanning calorimetry. Following a brief overview of the phase transition phenomenon exhibited by the lipid bilayer, we begin with the review by showing how several structural parameters underlying various phospholipids including phosphatidylcholine, phosphatidylethanolamine, and phosphatidylglycerol are defined and determined. Specifically, these structural parameters are obtained with saturated lipids packed in the gel-state bilayer using computer-based molecular mechanics calculations. Then we proceed to present the calorimetric data obtained with the lipid bilayer composed of saturated phospholipids as it undergoes the gel-to-liquid-crystalline phase transition in excess water. The general equations that can correlate the gel-to-liquid-crystalline phase transition temperature (T(m)) of the lipid bilayer with the structural parameters of the lipid molecule constituting the lipid bilayer are subsequently presented. From these equations, two tables of predicated T(m) values for well over 400 molecular species of saturated phosphatidylcholine and saturated phosphatidylethanolamine are generated. We further review the structure and chain-melting behavior of a large number of sn-1 saturated/sn-2 unsaturated phospholipids. Two T(m)-diagrams are shown, from which the effects of the number and the position of one to five cis carbon-carbon double bonds on T(m) can be viewed simultaneously. Finally, in the last part of this review, simple molecular models that have been invoked to interpret the characteristic T(m) trends exhibited by lipid bilayers composed of unsaturated lipids with different numbers and positions of cis carbon-carbon double bonds as seen in the T(m)-diagram are presented.  相似文献   

15.
Utilizing reflectance spectrophotometry, hemoperfusion, rate of oxygen consumption and redox level of mitochondrial cytochrome c (+c1) in livers in situ of anesthetized rats were measured. The transition to the anoxic state was induced by raising the pressure on the liver surface to more than the hepatic blood pressure by pressing with the tip of the optical guide of the reflectance spectrophotometer. During this transition, the average oxygen saturation of hemoglobin in the liver in situ decreased linearly with time until it became 10--20% of the total. This was followed by reduction of mitochondrial cytochrome c (+c1), which reached completion in 10--20 s. The measured O2 consumption rate remained constant until the percentage of oxyhemoglobin in situ decreased to a critical level. There was then a decrease in the rate of O2 consumption which was accompanied by a progressive reduction of cytochrome c (+c1). It was shown that amounts of hemoglobin and mitochondrial respiratory chain cytochromes in the liver in situ could be measured non-invasively and could provide important signals for vital cellular functions. The changes in hemoperfusion and rate of O2 consumption of the liver in situ following ethanol ingestion were also shown in rats, and are briefly discussed with respect to possible application of this method to study the pathophysiology of tissues.  相似文献   

16.
The hydrolysis of d-erythro beef brain sphingomyelin and d,l-erythro-N-palmitoylsphingomyelin dispersed as multilamellar liposomes by sphingomyelinase of Staphylococcus aureus is correlated with the thermotropic behavior of the sphingomyelins. In both cases maximal enzymatic hydrolysis was achieved at the beginning of the gel to liquid crystalline phase transition (30°C for beef brain sphingomyelin and 41°C for N-palmitoylsphingosinephosphorylcholine) with much lower activity both below and above these temperatures. The enzymatic activity was depressed in the presence of cholesterol in the bilayer which also depressed the phase transition. The profile of the enzymatic activity is explained by the uniqueness of the lipid molecules arrangement at the phase transition.  相似文献   

17.
After aqueous subcellular fractionation and partial purification by phosphocellulose chromatography, murine cells are found to contain a low molecular weight DNA-dependent DNA polymerase (beta) in the nuclear fraction and two distinguishable DNA-dependent DNA polymerases (C-I and C-II) in the cytosol. Both C-I and C-II are found in testis, liver, and regenerating liver; the amount of C-I being several fold increased in the regenerating liver and in immature testis. C-I and C-II are distinguishable by the criteria of salt sensitivity, inhibition by single-stranded DNA, elution from phosphocellulose, inhibition by 0.3 mM N-ethylmaleimide, template preference, and sedimentation coefficient. C-II is dissociated by 0.25 M KC1 to an active form of DNA polymerase of sedimentation coefficient 3.5 S while C-I is not dissociated, maintaining its sedimentation coefficient of 7.2 S. Many similar chemical and physical properties of C-II and the low molecular weight nuclear DNA polymerase (beta) suggest that C-II may represent an aggregate state of beta monomers, The size, reaction properties and the increase in enzyme activity under conditions of rapid cellular proliferation suggest C-I is analogous to the alpha DNA polymerase.  相似文献   

18.
Pure 1,2-dipalmitoyl-sn-glycero-3-phosphorylcholine (DPPC) or mixed DPPC:1,2-dipalmitoyl phosphatidyletanolamine (DPPE):1,2-dipalmitoyl diphosphatidylserine (DPPS) (17:5:3) liposomes were incorporated with 5 mol% dietary carotenoids (beta-carotene, lutein and zeaxanthin) or with cholesterol (16 and 48 mol%) in the absence or presence of 15 mol% carotenoids, respectively. The carotenoid incorporation yields ranged from 0.42 in pure to 0.72 in mixed phospholipid liposomes. They decreased significantly, from 3 to 14%, in the corresponding cholesterol-doped liposomes, respectively. Highest incorporation yields were achieved by zeaxanthin and lutein in phospholipid liposomes while in cholesterol-containing liposomes, lutein was highest incorporated. The effects on membrane structure and dynamics were determined by differential scanning calorimetry, steady-state fluorescence and anisotropy measurements. Polar carotenoids and cholesterol cause similar, dose-dependent effects: ordering and rigidification revealed by broadening of the transition peak, and increase of anisotropy. Membrane hydrophobicity is determined by cholesterol content and carotenoid polarity. In cholesterol-doped liposomes, beta-carotene is less incorporated than in cholesterol-free liposomes. Our observations suggest effects of carotenoids, even at much lower effective concentrations than cholesterol (8 to 80-fold), on membrane structure and dynamics. Although they are minor constituents of animal membranes, carotenoids may act as modulators of membrane phase transition, fluidity, polarity and permeability, and therefore, can influence the membrane physiology and pathology.  相似文献   

19.
The interaction of La3+ with phosphatidylserine vesicles is elucidated by binding studies, differential scanning calorimetry, X-ray diffraction, freeze fracture electron microscopy, and release of vesicle contents. La3+ effectively competes with Ca2+ for phosphatidylserine binding sites. The saturation level is close to a La/lipid ratio of 1:3. A concentration of 0.1 mM of La3+ is sufficient to induce fusion between sonicated vesicles.  相似文献   

20.
The interaction of La2+ with phosphatidylserine vesicles is studied by differential scanning calorimetry, 140La binding, 31P-NMR chemical shifts and relaxation rates, carboxyfluorescein and [14C]sucrose release, X-ray diffraction and freeze-fracture electron microscopy. In the presence of La3+ concentrations above 1 mM and an incubation temperature of 38 degrees C, i.e., at the phase transition temperature of the complex La/phosphatidylserine, the binding ratio of La/lipid exceeds a 1/3 ratio, reaching saturation at a 1/2 ratio. Analysis, employing a modified Gouy-Chapman equation, indicates a significant increase in the intrinsic binding constant of La/phosphatidylserine when the La3+ concentrations exceeds the threshold concentration for leakage. The analysis illustrates that at the molecular level the binding of La3+ can be comparable to or even weaker than that of Ca2+, but that even when present at smaller concentrations La3+ competes with and partially displaces Ca2+ from membranes or other negatively charged surfaces. The results suggest that the sequence La3+ greater than Ca2+ greater than Mg2+ reflects both the binding strength of these cations to phosphatidylserine as well as their ability to induce leakage, enhancement of 31P spin-lattice relaxation rates, fusion and other structural changes. The leakage, fusion, and other structural changes are more pronounced at the phase transition temperature of the La/lipid complex.  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