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.
Analysis of the time course of hydrolysis of dimyristoylphosphatidylcholine liposomes catalyzed by porcine pancreatic phospholipase A2 at 18 degrees C shows that, in the presence of 10 mM NaCl, the length of the latency period in the presteady-state phase increases from 3 to 10.5 min when the CaCl2 concentration is reduced from 15 to 1 mM. This inverse dependence of the lag period on calcium ion concentration is seen more readily at 1 M NaCl, where the induction time changes from 13.5 to 42 min by decreasing the concentration of CaCl2 from 15 to 1 mM. To interpret these results, we took into account the small amount of fatty acid that is produced during the latency phases. The fatty acid generates a negative surface electrostatic potential and makes the interfacial concentration of calcium ions different from the concentration in the bulk solvent. Variations in the analytical concentrations of NaCl and CaCl2 affect both the interfacial calcium ion concentration and electrostatic potential, as estimated theoretically from Grahame and Boltzmann equations. According to these estimates, the length of the latency period diminishes with the increase of the interfacial calcium concentration, but does not show any logical dependence on the change in surface electrostatic potential.  相似文献   

2.
D-Lactate dehydrogenase from the extreme halophilic archaebacterium Halobacterium marismortui has been partially purified by ammonium-sulfate fractionation, hydrophobic and ion exchange chromatography. Catalytic activity of the enzyme requires salt concentrations beyond 1M NaCl: optimum conditions are 4M NaCl or KCl, pH 6-8, 50 degrees C. Michaelis constants for NADH and pyruvate under optimum conditions of enzymatic activity are 0.070 and 4.5mM, respectively. As for other bacterial D-specific lactate dehydrogenases, fructose 1,6-bisphosphate and divalent cations (Mg2+, Mn2+) do not affect the catalytic activity of the enzyme. As shown by gel-filtration and ultracentrifugal analysis, the enzyme under the conditions of the enzyme assay is a dimer with a subunit molecular mass close to 36 kDa. At low salt concentrations (less than 1M), as well as high concentrations of chaotropic solvent components and low pH, the enzyme undergoes reversible deactivation, dissociation and denaturation. The temperature dependence of the enzymatic activity shows non-linear Arrhenius behavior with activation energies of the order of 90 and 25 kJ/mol at temperatures below and beyond ca. 30 degrees C. In the presence of high salt, the enzyme exhibits exceptional thermal stability; denaturation only occurs at temperatures beyond 55 degrees C. The half-time of deactivation at 70 and 75 degrees C is 300 and 15 min, respectively. Maximum stability is observed at pH 7.5-9.0.  相似文献   

3.
A number of breaks were recorded on the curve of Arrhenius relationship of the rate constant of the dye 1-anilino-8-naphthalenesulphonate sodium salt (ANS) input into human erythrocytes of 20, 28, 36, 42 and 46 degrees C. Variations in the values of activation energies within the temperature range of 28-36 degrees and 42-46 degrees C obtained in various blood samples allow to consider these temperatures as those at which structural changes of the membranes take place. The values of activation energy of the process for temperature "conformers" of the erythrocyte membrane are 12(10-20 degrees C), 26.5 (20-28 degrees C), 34.2(36-42 degrees C) and 47 kcal/mol (t is greater than 46 degrees C). Within the temperature range of 28-36 degrees and 42-46 degrees C an irreversible decrease of permeability to ANS of the erythrocyte ghost after their incubation for 10 min at increased temperatures were observed. Thus the temperature regions of the change in erythrocyte permeability correspond to those at which the resealing of ghost takes place. The break in Arrhenius graph at 20 degrees C seems to characterize a highly cooperative "point" transition. The lipid nature of the initiator of structural transition within 28-36 degrees C is proved by a sharp increase of the permeability of liposomes prepared from erythrocyte membrane lipids to ANS at 28 degrees C. The nature of the initiators of two other thermal transitions is discussed.  相似文献   

4.
An unusual effect of temperature on the ATPase activity of E. coli F1Fo ATP synthase has been investigated. The rate of ATP hydrolysis by the isolated enzyme, previously kept on ice, showed a lag phase when measured at 15 degrees C, but not at 37 degrees C. A pre-incubation of the enzyme at room temperature for 5 min completely eliminated the lag phase, and resulted in a higher steady-state rate. Similar results were obtained using the isolated enzyme after incorporation into liposomes. The initial rates of ATP-dependent proton translocation, as measured by 9-amino-6-chloro-2-methoxyacridine (ACMA) fluorescence quenching, at 15 degrees C also varied according to the pre-incubation temperature. The relationship between this temperature-dependent pattern of enzyme activity, termed thermohysteresis, and pre-incubation with other agents was examined. Pre-incubation of membrane vesicles with azide and Mg2+, without exogenous ADP, resulted in almost complete inhibition of the initial rate of ATPase when assayed at 10 degrees C, but had little effect at 37 degrees C. Rates of ATP synthesis following this pre-incubation were not affected at any temperature. Azide inhibition of ATP hydrolysis by the isolated enzyme was reduced when an ATP-regenerating system was used. A gradual reactivation of azide-blocked enzyme was slowed down by the presence of phosphate in the reaction medium. The well-known Mg2+ inhibition of ATP hydrolysis was shown to be greatly enhanced at 15 degrees C relative to at 37 degrees C. The results suggest that thermohysteresis is a consequence of an inactive form of the enzyme that is stabilized by the binding of inhibitory Mg-ADP.  相似文献   

5.
The phase behaviour of liposomes of 1,2-dimyristoyl-sn-glycero-3-phosphatidyl-sn-1'-glycerol (1'-DMPG) and the corresponding sn-3' stereoisomer (3'-DMPG) were studied by DSC as a function of NaCl concentration. The melting of the metastable gel phase to the liquid-crystalline phase was similar for both lipids. However, in the presence of salt and at 6 degrees C (T less than Tp) the gel phase of both stereoisomers of DMPG was shown to be metastable and a new phase nominated here as the highly crystalline phase was formed as the stable state. However, significant differences in the formation and melting of the highly crystalline phase were evident between the two polar headgroup stereoisomers. For 3'-DMPG in the presence of 300 mM NaCl the melting enthalpy of this phase is approx. 82 kJ/mol and the transition temperature about 11 degrees higher (at 33.6 degrees C) than for the gel to liquid-crystalline phase transition (25 kJ/mol at 23.0 degrees C). In the presence of 0.15-1.2 M NaCl at 6 to 10 degrees C the formation of the highly crystalline phase of 3'-DMPG is complete within 2 to 5 days, increasing [NaCl] facilitates the rate. For a 1:1 mixture of 1'- and 3'-DMPG the formation of the highly crystalline phase requires several weeks and melts at about 20 degrees higher than the gel phase (at approx. 40 degrees C). For 1'-DMPG partial conversion into the highly crystalline phase requires several months. For 3'-DMPG several intermediate phases appeared as endothermic peaks between the main phase transition temperature and the melting temperature of the highly crystalline phase. In contrast, for 1'-DMPG and the 1:1 mixture the subgel phase appears to be the only metastable intermediate phase. Different monovalent cations differ in their effect on the metastable behaviour.  相似文献   

6.
The hydrolysis of dipalmitoylphosphatidylcholine liposomes by porcine pancreatic phospholipase A2 was studied at 31 degrees C, i.e., with the substrate in the gel phase. Addition of delipidated bovine serum albumin to the assay medium induces the appearance of a latency phase in the time course of the enzymatic action. The lag period can be abolished by addition of free palmitic acid whereas no reversal by lysolecithin is found. The generation of a latency period by albumin appears to be due to its ability to sequester the palmitic acid newly released by the phospholipase A2 catalysis. Thus, the nascent fatty acid seems to be an essential activator of the enzymatic process.  相似文献   

7.
In the microsomal fraction of thyroid glands, the temperature dependence of DPH fluorescence polarization showed a discontinuity in the range of 29-33 degrees C. The transition temperatures of DMPC, DPPC and DSPC are near to the observed for the microsomal fraction. So that, thyroid peroxidase (TPO) was incorporated into liposomes made with these phospholipids. When DPH was incorporated in this peroxidase-liposome complex, a less pronounced phase transition was observed in the profiles of temperature dependence of DPH polarization, and the incorporation of the enzyme decreased the Tc. Arrhenius plots of TPO incorporated into liposomes showed discontinuities at similar temperatures observed by fluorescence polarization. The decrease of transition temperature of liposomes induced by thyroid peroxidase incorporation suggests that this enzyme seems to need a fluid medium for its enzyme activity.  相似文献   

8.
Sphingomonas sp. strain Ant 17 was isolated from fuel-contaminated soil collected at Scott Base, Ross Island, Antarctica. We anticipated that Ant 17 would be a good model organism for studying cold climate bioremediation, and therefore determined its biodegradation capabilities and tolerance of potentially growth-limiting environmental conditions. Sphingomonas sp. Ant 17 degrades the aromatic fraction of several different crude oils, jet fuel, and diesel fuel at low temperatures and without nutrient amendment. It utilizes or transforms a broad range of pure aromatic substrates, including hydrocarbons, heterocycles, and aromatic acids and alcohols. Ant 17 grows at temperatures of 1 degree C to 35 degrees C and mineralizes radiolabeled phenanthrene over a range of more than 24 degrees C. This psychrotolerant isolate appears to utilize hydrocarbons more efficiently at low temperatures than would be predicted by mesophilic enzyme kinetics. The optimum pH for growth was 6.4 at 22 degrees C, with extended lag phases observed in more alkaline media. However, there was less effect of pH on lag phase at lower temperatures. Ant 17 displayed greater tolerance to UV irradiation and freeze-thaw cycles than the hydrocarbon-degrading isolate Sphingomonas sp. WPO-1, which may reflect adaptation to its Antarctic soil environment. However, it was more sensitive than expected to desiccation and to low concentrations of NaCl and CaCl(2). Ant 17 was phenotypically stable and lacked detectable plasmids, suggesting a chromosomal location for genes encoding aromatic degradation enzymes. Its broad aromatic substrate range and tolerance of low and fluctuating temperature and low nutrients make Sphingomonas sp. Ant 17 a valuable microbe for examining fuel spill bioremediation in cold soils.  相似文献   

9.
The soluble Ca2+-ATPase from Escherichia coli had a distinctive behavior with respect to inhibition by Na+ measured at 36 degrees C and 19 degrees C. At the first temperature the Hill plots are linear and show a slope of 1.1 while at 19 degrees C the plots are biphasic, with slopes of 1.8 and 0.8 before and after the break, respectively. The break occurs at about 50 nM NaCl. Gel chromatography was performed in jacketed Sepharose 4B columns kept at 2 temperatures in the presence of different concentrations of NaCl. It was found that the Stokes' radius of the enzyme was dependent on the temperature and on the salt concentration. Equilibrium sucrose gradients run at 19 degrees C showed that the sedimentation constant of the enzyme remained constant irrespective of the NaCl concentration used. It is concluded that a "folding" of the enzyme takes place in the presence of NaCl, the process being complete at about 50 mM NaCl at 19 degrees C and at about 20 mM at 36 degrees C. The results are in excellent agreement with the kinetic data: the "folded" or "compact" configuration would show no cooperative response towards Na+ while the "expanded" conformer would present strong cooperativity. This is also in agreement with the results obtained with the enzyme embedded in the membrane: when the membrane is fluid a high n value (Hill coefficient) is found; when the membrane is more rigid the value of n falls. A model explaining all our results is proposed and discussed.  相似文献   

10.
The binding of labeled hyaluronate to the surface of Simian virus 40-transformed 3T3 cells was studied as a function of 1) pH, 2) ionic strength, 3) temperature, and 4) molecular weight of the hyaluronate. Binding occurred over a wide range of pH values with optima at pH 7 and at less than pH 4. Binding at low pH was eliminated at high ionic strength whereas that at physiological pH was enhanced, with a maximum at 0.5 M NaCl. The enhancement of binding at pH 7 was reversible and independent of the particular salt used. Scatchard plot analysis showed that increasing the ionic strength resulted in both a decrease in the dissociation constant (Kd) and an increase in the amount bound at saturation (Bmax). Temperature also influenced the binding of hyaluronate to the cell surface. The amount bound at low temperatures (0 degrees C) was 3 to 5 times that bound at high temperatures (40 degrees C) with a sharp transition occurring at 18 degrees C, the temperature of phase transition of the plasma membrane. The temperature effect was primarily a change in the Bmax and was reversible. Finally the molecular weight of the ligand influenced the binding. High molecular weight preparations of hyaluronate had a higher binding affinity (lower Kd) and a lower Bmax than did smaller molecular weight preparations.  相似文献   

11.
12.
We have studied the binding of CTP: phosphocholine cytidylyltransferase from HeLa cell cytosol to large unilamellar vesicles of egg phosphatidylcholine (PC) or HeLa cell phospholipids that contain various amounts of oleic acid. A fatty acid/phospholipid molar ratio exceeding 10% was required for CTP: phosphocholine cytidylyltransferase binding to liposomes. At a fatty acid/phospholipid molar ratio of 1; 85% of the cytosolic CTP: phosphocholine cytidylyltransferase was bound. The enzyme also bound to liposomes with at least 20 mol% palmitic acid, monoolein, diolein or oleoylacetylglycerol. Oleoyl-CoA did not promote enzyme binding to liposomes. Binding to oleate-PC vesicles was blocked by Triton X-100 but not by 1 M KCl, and was reversed by incubation of the vesicles with bovine serum albumin. Cytidylyltransferase bound to egg PC vesicles that contained 33 mol% oleic acid equally well at 4 degrees C and 37 degrees C. The enzyme also bound to dimyristoyl- and dipalmitoylphosphatidylcholine vesicles containing oleic acid at temperatures below the phase transition for these liposomes. Binding of the cytidylyltransferase to egg PC vesicles containing oleic acid, monoolein, oleoylacetylglycerol or diolein resulted in enzyme activation, as did binding to dipalmitoylPC-oleic acid vesicles. However, binding to egg PC-palmitic acid vesicles did not fully activate the transferase. Various mechanisms for cytidylyltransferase interaction with membranes are discussed.  相似文献   

13.
Molecular packing and the thermotropic phase behavior of fully hydrated ammonium salts of 1,2-dimyristoyl-sn-glycero-3-phosphatidyl-sn-1'-glycerol (1'-DMPG) and the corresponding 3' stereoisomer (3'-DMPG) as well as the effects of 300 mM NaCl on these lipids were studied by Fourier transform infrared (FTIR) spectroscopy. The ammonium salts of both stereoisomer show similar thermotropic phase behavior and have an order-disorder phase transition at approximately 21 degrees C. While complexing with Na+, however, an incubation of liposomes at +6 degrees C for 3 days results in significant structural differences between liposomes of 1'-DMPG and 3'-DMPG. In the presence of 300 mM NaCl the infrared spectra for 3'-DMPG reveal the appearance of a more solidified lipid nominated here as the highly crystalline phase with a transition into the liquid-crystalline state at a significantly higher temperature (approximately at 33 degrees C) than that for 1'-DMPG (approximately at 23 degrees C). Crystal field splitting resulting from interchain vibrational coupling is observed in the CH2 scissoring mode of the 3'-DMPG(Na+) complex in the highly crystalline phase (T less than 33 degrees C); i.e., the acyl chains are packed in a rigid orthorhombic- or monoclinic-like crystal lattice. At temperatures above the transition at 33 degrees C the acyl chains of 3'-DMPG(Na+) give rise to infrared spectra indicative of hexagonal packing. The latter type of hydrocarbon chain packing is also found for the ammonium salts of 1'-DMPG and 3'-DMPG without Na+ as well as for 1'-DMPG with Na+. In addition, the binding of Na+ to 3'-DMPG causes narrowing of the bands associated with the interfacial and polar headgroup regions of 3'-DMPG and thus reveals reduced motional freedom. This demonstrates that Na+ binds tightly to 3'-DMPG, leading to the immobilization of the entire phospholipid polar headgroup. Such effects by Na+ are not observed for 1'-DMPG.  相似文献   

14.
A phospholipid analog 1-palmitoyl-2-6(pyren-1-yl)hexanoyl-sn-glycero-3-phospho-N- (trinitrophenyl)aminoethanol (PPHTE) in which pyrene fluorescence is intramolecularly quenched by the trinitrophenyl group was used as a substrate for pancreatic phospholipase A2. Upon phospholipase A2 catalyzed hydrolysis of this molecule pyrene monomer fluorescence emission intensity increased as a result of the transfer of the pyrene fatty acid to the aqueous phase. Optimal conditions for phospholipase A2 hydrolysis of PPHTE were similar to those observed earlier for other pyrenephospholipids (T. Thuren, J. A. Virtanen, R. Verger, and P. K. J. Kinnunen (1987) Biochim. Biophys. Acta 917, 411-417). Although differential scanning calorimetry revealed no thermal phase transitions for PPHTE between +5 and +60 degrees C the Arrhenius plot of the enzymatic hydrolysis of the lipid showed a discontinuity at 30 degrees C. The molecular origin of this discontinuity remains at present unknown. To study the effects of dimyristoylphosphatidylcholine (DMPC) phase transition at 23.9 degrees C on phospholipase A2 reaction PPHTE was mixed with DMPC in a molar ratio of 1:200 in small unilamellar vesicles. The hydrolysis of DMPC-PPHTE vesicles was measured by following the increase in pyrene monomer fluorescence emission due to phospholipase A2 action on PPHTE. Below the phase transition of DMPC the enzymatic reaction exhibited a hyperbolic behavior. At the transition as well as at slightly higher temperatures a lag period was observed. The longest lag period was approximately 20 min. Above 26 degrees C no lag time could be observed. However, the reaction rates were slower than below the phase transition temperature.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
The interaction of aqueous phospholipid dispersions of negatively charged 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol, sodium salt (DMPG) with the divalent cations Mg(2+), Ca(2+) and Sr(2+) at equimolar ratios in 100 mM NaCl at pH 7 was investigated by Fourier transform infrared spectroscopy. The binding of the three cations induces a crystalline-like gel phase with highly ordered and rigid all-trans acyl chains. These features are observed after storage below room temperature for 24 h. When the gel phase is heated after prolonged incubation at low temperature phase transitions into the liquid crystalline phase are observed at 58 degrees C for the DMPG:Sr(2+), 65 degrees C for the DMPG:Mg(2+), and 80 degrees C for the DMPG:Ca(2+) complex. By subsequent cooling from temperatures above T(m) these complexes retain the features of a liquid crystalline phase with disordered acyl chains until a metastable gel phase is formed at temperatures between 38 and 32 degrees C. This phase is characterized by predominantly all-trans acyl chains, arranged in a loosely packed hexagonal or distorted hexagonal subcell lattice. Reheating the DMPG:Sr(2+) samples after a storage time of 2 h at 4 degrees C results in the transition of the metastable gel to the liquid crystalline phase at 35 degrees C. This phase transition into the liquid crystalline state at 35 degrees C is also observed for the Mg(2+) complex. However, for DMPG:Mg(2+) at higher temperatures, a partial recrystallization of the acyl chains occurs and the high temperature phase transition at 65 degrees C is also detected. In contrast, DMPG:Ca(2+) exhibits only the phase transition at 80 degrees C from the crystalline gel into the fluid state upon reheating. Below 20 degrees C, the rate of conversion from the metastable gel to a thermodynamically stable, crystalline-like gel phase decreases in the order Ca(2+)&z. Gt;Mg(2+)>Sr(2+). This conversion into the crystalline gel phase is accompanied by a complete dehydration of the phosphate groups in DMPG:Mg(2+) and by a reorientation of the polar lipid head groups in DMPG:Ca(2+) and in DMPG:Sr(2+). The primary binding sites of the cations are the PO(2)(-) groups of the phosphodiester moiety. Our infrared spectroscopic results suggest a deep penetration of the divalent cations into the polar head group region of DMPG bilayers, whereby the ester carbonyl groups, located in the interfacial region of the bilayers, are indirectly affected by strong hydrogen bonding of immobilized water molecules. In the liquid crystalline phase, the interaction of all three cations with DMPG is weak, but still observable in the infrared spectra of the DMPG:Ca(2+) complex by a slight ordering effect induced in the acyl chains, when compared to pure DMPG liposomes.  相似文献   

16.
W F Stafford 《Biochemistry》1985,24(13):3314-3321
The stability of skeletal myosin rod was studied by following the dependence of both papain digestion kinetics and helix-coil transition temperatures on the concentration of neutral salts. The rate of papain-catalyzed digestion of rod to form subfragment 2 and light meromyosin was strongly dependent on chloride concentration but essentially independent of acetate concentration up to 2.0 M. The rod exhibited a biphasic melting curve in 0.6 M NaCl, 5 mM phosphate, and 0.1 mM ethylenediaminetetraacetic acid (EDTA), pH 7.3, with transitions at 45 and 53 degrees C. In 0.6 M CH3COONa, 5 mM phosphate, and 0.1 mM EDTA, pH 7.3, the transitions occurred at 50 and 58 degrees C, respectively. Transition temperatures were obtained with a novel curve-fitting method. The effect of increasing chloride ion concentration on melting profiles was 2-fold. Below 0.6 M salt, the two transition temperatures, Tm,1 and Tm,2, depended on salt concentration such that increasing NaCl concentration caused a small stabilization of the helix while increasing acetate concentration caused the helix to become markedly more stable. Between 0.6 and 1.0 M, variation of chloride concentration had almost no effect on the thermal stability of the rod while increasing acetate concentration increased its stability considerably. Above 1.0 M NaCl, the melting profiles became broad with a third transition being observed (e.g., at 3.0 M, Tm,3 = 38 degrees C), indicating the existence of a region which has a tendency to be destabilized by chloride. The third transition was not observed at comparable concentrations of acetate. This effect of chloride was not expected on the basis of its position in the Hofmeister series.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The effect of ions on the thermostability and unfolding of Na,K-ATPase from shark salt gland was studied and compared with that of Na,K-ATPase from pig kidney by using differential scanning calorimetry (DSC) and activity assays. In 1 mM histidine at pH 7, the shark enzyme inactivates rapidly at 20 degrees C, as does the kidney enzyme at 42 degrees C (but not at 20 degrees C). Increasing ionic strength by addition of 20 mM histidine, or of 1 mM NaCl or KCl, protects both enzymes against this rapid inactivation. As detected by DSC, the shark enzyme undergoes thermal unfolding at lower temperature (Tm approximately 45 degrees C) than does the kidney enzyme (Tm approximately 55 degrees C). Both calorimetric endotherms indicate multi-step unfolding, probably associated with different cooperative domains. Whereas the overall heat of unfolding is similar for the kidney enzyme in either 1 mM or 20 mM histidine, components with high mid-point temperatures are lost from the unfolding transition of the shark enzyme in 1 mM histidine, relative to that in 20 mM histidine. This is attributed to partial unfolding of the enzyme due to a high hydrostatic pressure during centrifugation of DSC samples at low ionic strength, which correlates with inactivation measurements. Addition of 10 mM NaCl to shark enzyme in 1 mM histidine protects against inactivation during centrifugation of the DSC sample, but incubation for 1 h at 20 degrees C prior to addition of NaCl results in loss of components with lower mid-point temperatures within the unfolding transition. Cations at millimolar concentration therefore afford at least two distinct modes of stabilization, likely affecting separate cooperative domains. The different thermal stabilities and denaturation temperatures of the two Na,K-ATPases correlate with the respective physiological temperatures, and may be attributed to the different lipid environments.  相似文献   

18.
Large unilamellar dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylglycerol (DPPG) liposomes loaded with an aqueous chemotherapeutic drug, cytosine arabinofuranoside (ARA-C), were exposed for 30 min to 60 W/kg continuous-wave (CW) 100-MHz or 2.45-GHz radiation in vitro at temperatures between 37 degrees C and 43 degrees C. Liposomes were exposed in HEPES buffer or in HEPES buffer supplemented with 44% by volume fetal calf serum (FCS). Characteristic phase transition responses were detected in the range of 39 degrees C to 40 degrees C with the presence of FCS, increasing maximum % release of 3H-ARA-C by 20% relative to HEPES suspension. Neither frequency of electromagnetic radiation had any detectable effect on liposome permeability or the location of the phase transition in the presence or absence of FCS.  相似文献   

19.
The influence of sucrose, between 0 and 70% in the aqueous phase, upon multilamellar liposomes of dimyristoylphosphatidylcholine was examined by differential scanning calorimetry and X-ray diffraction analysis. Increasing concentrations of sucrose increase the temperatures of both the main transition and the pretransition of the lipid. The effect is greater on the pretransition than on the main transition. At 35 degrees C the interlamellar spacing in the multilamellar liposomes is reduced by increasing sucrose concentration in the aqueous phase and no significant effects are seen in the chain lattice of the bilayers. This result is interpreted as a dehydrating effect of sucrose upon the bilayer-water system at 35 degrees C. At 5 degrees C the interlamellar spacing is increased and this increase is, at high (70%) sucrose concentrations, attributable to an untilting of the lipid acyl chains with no change in the thickness of the aqueous layers in the multilamellae.  相似文献   

20.
G Cevc  A Watts  D Marsh 《Biochemistry》1981,20(17):4955-4965
The dependence of the gel-to-fluid phase transition temperature of dimyristoyl- and dipalmitoylphosphatidylserine bilayers on pH, NaCl concentration, and degree of hydration has been studied with differential scanning calorimetry and with spin-labels. On protonation of the carboxyl group (pK2app = 5.5), the transition temperature increases from 36 to 44 degrees C in the fully hydrated state of dimyristoylphosphatidylserine (from 54 to 62 degrees C for dipalmitoylphosphatidylserine), at ionic strength J = 0.1. In addition, at least two less hydrated states, differing progressively by 1 H2O/PS, are observed at low pH with transition temperatures of 48 and 52 degrees C for dimyristoyl- and 65 and 68.5 degrees C for dipalmitoylphosphatidylserine. On deprotonation of the amino group (pK3app = 11.55) the transition temperature decreases to approximately 15 degrees C for dimyristoyl- and 32 degrees C for dipalmitoylphosphatidylserine, and a pretransition is observed at approximately 6 degrees C (dimyristoylphosphatidylserine) and 21.5 degrees C (dipalmitoylphosphatidylserine), at J = 0.1. No titration of the transition is observed for the fully hydrated phosphate group down to pH less than or equal to 0.5, but it affinity for water binding decreases steeply at pH greater than or equal to 2.6. Increasing the NaCl concentration from 0.1 to 2.0 M increases the transition temperature of dimyristoyphosphatidylserine by approximately 8 degrees C at pH 7, by approximately 5 degrees at pH 13, and by approximately 0 degrees C at pH 1. These increases are attributed to the screening of the electrostatic titration-induced shifts in transition temperature. On a further increase of the NaCl concentration to 5.5 M, the transition temperature increases by an additional 9 degree C at pH 7, 13 degree C at pH 13, approximately 7 degree C in the fully hydrated state at pH 1, and approximately 4 and approximately 0 degree C in the two less hydrated states. These shifts are attributed to displacement of water of hydration by ion binding. From the salt dependence it is deduced that the transition temperature shift at the carboxyl titration can be accounted for completely by the surface charge and change in hydration of approximately 1 H2O/lipid, whereas that of the amino group titration arises mostly from other sources, probably hydrogen bonding. The shifts in pK (delta pK2 = 2.85, delta pK3 = 1.56) are consistent with a reduced polarity in the head-group region, corresponding to an effective dielectric constant epsilon approximately or equal to 30, together with surface potentials of psi congruent to -100 and -150 mV at the carboxyl and amino group pKs, respectively. The transition temperature of dimyristoylphosphatidylserine-water mixtures decreases by approximately 4 degree C each water/lipid molecule added, reaching a limiting value at a water content of approximately 9-10 H2O/lipid molecule.  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