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.
M Caffrey 《Biochemistry》1985,24(18):4826-4844
A study of the kinetics and mechanism of the thermotropic lamellar gel/lamellar liquid-crystalline and lamellar/inverted hexagonal phase transition in dihexadecylphosphatidylethanolamine (DHPE) at various hydration levels has been carried out. Measurements were made by using a real-time X-ray diffraction method at the Cornell High Energy Synchrotron Source. This represents an extension of an earlier study concerning the lamellar gel/lamellar liquid-crystalline phase transition in dipalmitoylphosphatidylcholine [Caffrey, M., & Bilderback, D. H. (1984) Biophys. J. 45, 627-631]. With DHPE, the chain-melting and the nonbilayer transitions were examined under active heating and passive cooling conditions by using a temperature jump to effect phase transformation. Measurements were made at hydration levels ranging from 0% to 60% (w/w) water, and in all cases, the transitions were found to be repeatable, be reversible, and have an upper bound on the transit times (time required to complete the transition) of less than or equal to 3 s. The shortest transit time recorded for the chain-melting and lamellar/hexagonal transitions was less than 1 s. At 8% (w/w) water, the transit times were still on the order of seconds even though the transition does not involve the intermediate L alpha phase. Note, the measured transit times are gross values incorporating the intrinsic transit time in addition to the time required to heat or cool the sample through the transition temperature range and to supply or remove the latent heat of the transition. Regardless of the direction of the transition, both appear to be two state to within the sensitivity limits of the real-time method. From simultaneous wide- and low-angle measurements at the lamellar chain-melting transition, loss of long-range order in the lamellar gel phase appears to precede the chain-melting process. On the basis of the real-time X-ray diffraction measurements, a mechanism is proposed for the lamellar/hexagonal phase transition. The mechanism does not involve large or energetically expensive molecular rearrangements, leads directly to a hexagonal lattice coplanar with the lamellar phase, incorporates facile reversibility, repeatability, and cooperativity, accounts for an observed, apparent memory in the hexagonal phase of the original lamellar phase orientation, and is consistent with the experimental observation of a predominantly two-state transition. In conjunction with the kinetic measurements, the DHPE/water phase diagram was constructed. At and above 12% (w/w) water, the thermotropic transition sequence is L beta'/L alpha/HII.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Helix-coil transition of DNA with attached extended ligands able to interact with one another during adsorption on DNA (cooperative or uncooperative binding) has been considered. The general formulae describing dependence of polymer melting curve on concentration of attached ligands have been obtained. It has been shown that cooperativity of interaction with DNA stipulates for two phase profile of the melting curve. The results obtained show that proteins which interact with DNA cooperatively may cause two phase helix-coil transition under conditions of reversible binding.  相似文献   

3.
Although sugar has been suggested to promote floral transition in many plant species, growth on high concentrations (5% [w/v]) of sucrose (Suc) significantly delayed flowering time, causing an increase in the number of leaves at the time of flowering in Arabidopsis. The effect of high concentrations of Suc seemed to be metabolic rather than osmotic. The delay of floral transition was due to extension of the late vegetative phase, which resulted in a delayed activation of LFY expression. In addition, growth on low concentrations (1% [w/v]) of Suc slightly inhibited flowering in wild-type plants. This delay resulted from effects on the early vegetative phase. This inhibition was more pronounced in tfl1, an early flowering mutant, than in the wild type. Although 1% (w/v) Suc was reported to promote floral transition of late-flowering mutants such as co, fca, and gi, floral transition in these mutants was delayed by a further increase in Suc concentration. These results suggest that sugar may affect floral transition by activating or inhibiting genes that act to control floral transition, depending on the concentration of sugars, the genetic background of the plants, and when the sugar is introduced. Growth on 1% (w/v) Suc did not restore the reduced expression levels of FT and SOC1/AGL20 in co or fca mutants. Rather, expression of FT and SOC1/AGL20 was repressed by 1% (w/v) Suc in wild-type background. The possible effects of sugar on gene expression to promote floral transition are discussed.  相似文献   

4.
The ultrasonic absorption coefficient per wavelength (alpha lambda), as a function of temperature and frequency, was determined for large unilamellar vesicles (LUV) in the vicinity of their phospholipid phase transition temperature, using a double crystal acoustic interferometer. (The vesicles were composed of a 4:1 (w/w) mixture of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylglycerol (DPPG). It has been found that alpha lambda reaches a maximum (alpha lambda)max at the phase transition temperature (tm) of the phospholipids in the bilayer, at an ultrasonic relaxation frequency of 2.1 MHz. Divalent cations (Ca2+ and Mg2+), added to LUV suspensions, shifted (alpha lambda)max to higher temperatures, dependent upon the concentration of divalent cation. It was also found that the shape of the alpha lambda versus t curve was significantly changed, representing changes in the Van't Hoff enthalpy of the transition, and therefore, the cooperative unit of the transition. This suggests that divalent cations interact individually with the negatively charged phospholipid headgroups of DPPG and with DPPC headgroups, thus decreasing the cooperative unit of the transition. The observed upward shift in tm suggests an interaction that increases the activation energy and, therefore, the temperature of the phase transition. However, alpha lambda as a function of frequency did not change with the addition of divalent cations and, thus, the relaxation time of the event responsible for the absorption of ultrasound is not changed by the addition of divalent cations.  相似文献   

5.
Myelin proteolipid has been isolated from bovine brain and purified using organic solvents according to conventional procedures. The protein content of the purified sample, or crude proteolipid, contains a minimum of 75% w/w of proteolipid, with DM-20, a proteolipid molecule with an internal deletion of 35 out of 276 amino acid residues, as the only other component. Biochemical analysis has shown the differences in lipid composition between brain white matter, myelin and crude proteolipid preparations. The latter contained practically no cholesterol, while the other two samples had about 22-23% w/w. High-sensitivity differential scanning calorimetry experiments with both crude proteolipid and its extracted pool of lipids have shown similar reversible thermal transitions at 52 degrees C and 48 degrees C. The effect of increasing amounts of cholesterol on the two calorimetric transitions led in both cases to a continuous decrease in the melting temperature and in the transition enthalpy. Parallel Fourier-transform infrared spectroscopy studies of crude proteolipid have detected a reversible, co-operative lipid transition centred at 49 degrees C, with no detectable change in the amide region between 20 degrees C and 60 degrees C. Once more an increase in cholesterol content led to a decrease in the sharpness of this transition. It is concluded that the thermal transition detected in crude proteolipid, which has in the past been attributed to proteolipid thermal denaturation (Mateo et al. 1986), actually corresponds to a thermotropic phase transition of the lipids included in the crude proteolipid sample.  相似文献   

6.
Neutral Lipids Rigidify Unsaturated Acyl Chains in Senescing Membranes   总被引:3,自引:0,他引:3  
Senescence in bean cotyledons is accompanied by a progressiveincrease in the proportion of gel phase lipid in cellular membranesthat can be attributed to qualitative changes in the neutrallipids. The resulting mixture of lipid phases leads to impairedmembrane function. Insight into the molecular basis for thisphenomenon has been gleaned from studies of the effects of theseneutral lipids on the phase properties of pure phospholipidmembranes. Induction of the gel phase, detectable as a risein the liquid-crystalline to gel phase transition temperature,was observed when neutral lipid from senescent membranes wasintroduced into liposomes of unsaturated phospholipids. Thetransition temperature for phosphatidylcholine rose from –6.5° C in control liposomes to 51 ° C when 25% (w/w) neutrallipid was present. A similar rise was obtained for dioleoylphosphatidylcholine.However, for dipalmitoylphosphatidylcholine, which is fullysaturated, the rise in transition temperature upon additionof neutral lipids was only 3 ° C. Thus the neutral lipidsin senescent membranes appear to selectively rigidify unsaturatedacyl chains. At least three types of compounds known to alterthe phase properties of lipid bilayers were detectable in theneutral lipid fraction, suggesting that rigidification reflectsthe concerted action of several neutral lipid components onunsaturated phospholipid. Key words: Senescence, phospholipid, lipid phase properties  相似文献   

7.
1-Behenyl-2-lauryl-sn-glycero-3-phosphocholine (22/12 PC) belongs to a unique group of phospholipids in which the molecule has one acyl chain almost twice as long as the other. The temperature-composition phase diagram for this lipid in the range of 25-65 degrees C, and 0 to 84.3% (w/w) water has been constructed by using the isoplethal method in the heating direction and x-ray diffraction for phase identification and structure characterization. At water contents between 10.3 and 34% (w/w) and at temperatures below 43 degrees C, a single mixed interdigitated lamellar gel phase (Lm beta, [symbol: see text]) of the type described by Hui et al. (1984. Biochemistry. 23:5570-5577) and McIntosh et al. (1984. Biochemistry. 23:4038-4044) was found. A second phase consisting of bulk aqueous solution coexists with the Lm beta phase at hydration levels above 34% (w/w) water in the temperature range between 25 and 43 degrees C. Above 43 degrees C, a partially interdigitated lamellar liquid crystalline (Lp alpha) phase ([symbol: see text]) is seen in the water concentration range extending from 0 to 84.3% (w/w). The pure Lp alpha phase is found below 43% (w/w) water, while coexistence of the Lp alpha phase and the bulk aqueous solution is observed above this water concentration which marks the hydration boundary. Interestingly, the latter boundary for both Lm beta and Lp alpha phases is nearly vertical in the temperature range studied. Furthermore, the lamellar chain-melting transition temperature appears to be relatively insensitive to hydration in the range 0-85% (w/w) water. We have confirmed the identify of the Lm beta phase by constructing a 5.7-A resolution electron density profile on oriented samples by the swelling method. Temperature-induced chain melting effects an increase in lipid bilayer thickness suggesting that the Lp alpha phase has chains packed in the partially as opposed to the mixed interdigitated configuration. Unlike the symmetric phosphatidylcholines a ripple (P beta') phase was not found as an intermediate between the low and high temperature lamellar phases of 22/12 PC. The specific volume of 22/12 PC is 940 (+/- 1) microliter/g and 946 (+/- 1) microliter/g in the hydrated lamellar gel state at 28 (+/- 2) and 40 (+/- 2) degrees C, respectively, from neutral buoyancy experiments. Based on measurements of the temperature dependence of the various lattice parameters of the different phases encountered in this study the corresponding lattice thermal expansion coefficients have been measured. These are discussed and their dependence on lipid hydration is reported.  相似文献   

8.
Aqueous dispersions of a porcine lung surfactant (PLS) extract with and without cholesterol supplementation were analyzed by X-ray scattering. Lamellar liquid-crystalline and gel-type bilayer phases are formed, as in pure phosphatidylcholine (PC)-cholesterol systems. This PLS extract, developed for clinical applications, has a cholesterol content of less than 1% (w/w). Above the limit of swelling, the bilayer structure shows a melting (main) transition during heating at about 34 degrees C. When 13 mol% cholesterol was added to PLS, so that the cholesterol content of natural lung surfactant was reached, the X-ray scattering pattern showed pronounced changes. The main transition temperature was reduced to the range 20-25 degrees C, whereas according to earlier studies of disaturated PC-cholesterol bilayers in water the main transition remains almost constant when the amount of solubilized cholesterol is increased. Furthermore, the changes in scattering pattern at passing this transition in PLS-cholesterol samples were much smaller than at the same transition in PLS samples. These effects of cholesterol solubilization can be related to phase segregation within the bilayers, known from pure PC-cholesterol systems. One phase, solubilizing about 8 mol% cholesterol, exhibits a melting transition, whereas the other bilayer phase, with a liquid-crystalline disordered conformation, has a cholesterol content in the range 20-30 mol% and this phase shows no thermal transition. The relative amount of bilayer lipids that is transformed at the main transition in the PLS-cholesterol sample is therefore only half compared to that in PLS samples. The reduction in transition temperature in the segregated bilayer of lung surfactant lipids is probably an effect of enrichment of disaturated PC species in the phase, which is poor in cholesterol. This work indicates that cholesterol in lung surfactant regulates the crystallization behavior.  相似文献   

9.
The phase behaviour of leaf polar lipids from three plants, varying in their sensitivity to chilling, was investigated by differential scanning calorimetry. For the lipids from mung bean (Vigna radiata L. var. Berken), a chilling-sensitive plant, a transition exotherm was detected beginning at 10 ± 2°C. No exotherm was evident above 0°C with polar lipids from wheat (Triticum aestivum cv. Falcon) or pea (Pisum sativum cv. Massey Gem), plants which are insensitive to chilling. The enthalpy for the transition in the mung bean polar lipids indicated that only about 7% w/w of the lipid was in the gel phase at ?8°C. The thermal transition of the mung bean lipids was mimicked by wheat and pea polar lipids after the addition of 1 to 2% w/w of a relatively high melting-point lipid such as dipalmitoylphosphatidylcholine, dipalmitoylphosphatidylglycerol or dimyristoylphosphatidylcholine. Analysis of the polar lipids from the three plants showed that a dipalmitoylphosphatidylglycerol was present in mung bean (1.7% w/w) and pea (0.3% w/w) but undetected in wheat, indicating that the transition exotherm temperature of 10°C in mung bean, 0°C in pea and about ?3°C in wheat correlates with the proportion of the high melting-point disaturated component in the polar lipids. The results indicate that the transition exotherm, observed at temperatures above 0°C in the membranes of chilling-sensitive plants, could be induced by small amounts of high melting-point lipids and involves only a small proportion of the membrane polar lipids.  相似文献   

10.
Steady-state and time-resolved emission spectroscopy of 1-anilinonaphthalene-8-sulfonic acid (ANS) have been used for characterization of the metastable rippled gel phase, Pbeta'(mst), formed in fully-hydrated dipalmitoylphosphatidylcholine (DPPC) upon cooling from the liquid crystalline phase Lalpha [Tenchov et al., Biophys. J. 56 (1989) 757]. The Pbeta'(mst) phase of DPPC clearly differs from the stable Pbeta' phase by increased (approximately 27%) ANS emission intensity, by enhanced (approximately 23%) average radiative rate constant, and by reduced (approximately 18%) non-radiative quenching rate constant. The fluorescence intensity peak at the Pbeta'-->Lalpha transition temperature is replaced by a large, reversible stepwise intensity drop at the Pbeta'(mst)-->Lalpha transition. No such effects have been found for dimiristoylphosphatidylcholine (DMPC) dispersions confirming previous results that DMPC does not form a Pbeta'(mst) phase. Since ANS is known to predominantly reside in the interfacial region, the observed effects indicate differences between the stable and metastable rippled phases in the organization and dynamics of their lipid/water interfaces. The data demonstrate that the metastable rippled phase manifests its appearance also through interactions with small molecules (ANS size approximately 8 A).  相似文献   

11.
D H Croll  D M Small  J A Hamilton 《Biochemistry》1985,24(27):7971-7980
The phase behavior of cholesteryl esters with triglyceride has been characterized by differential scanning calorimetry (DSC), light microscopy, and polarizing light microscopy (PLM). Temperature-dependent molecular motions determined by 13C NMR spectroscopy were correlated with thermotropic phase behavior. Two systems, cholesteryl oleate (CO) and a 3/1 w/w mixture of cholesteryl linoleate (CL) and CO, were examined in the presence of small amounts of triolein (TO). Both systems exhibited metastable cholesteric and smectic (or only smectic) phases. Increasing amounts of TO progressively lowered the liquid-crystalline phase transition temperatures and eventually abolished the cholesteric phase, but at differing amounts of TO for the two systems (between 4% and 5% with CL/CO and between 7% and 10% with CO). DSC and PLM showed a progressive broadening of the phase transitions as well as an overlapping of the temperature ranges of the cholesteric and smectic phases. At greater than or equal to 4% TO, a separate isotropic liquid phase coexisted with liquid-crystalline phases. 13C NMR spectroscopy was used to monitor the molecular motions of the cholesteryl ester steroid ring and acyl chain in liquid and liquid-crystalline phases. In the liquid phase, no significant changes in fatty acyl motions, as reflected in spin-lattice relaxation time (T1) and nuclear Overhauser enhancement (NOE) values, were found on addition of TO. The line width (v 1/2) of the steroid ring resonances increased markedly near (1-5 degrees C above) the isotropic liquid----liquid-crystal phase transition temperature (TLC). However, the C3/C6 v 1/2 ratio at 1 degree C above TLC was greater for mixtures exhibiting an isotropic----cholesteric transition than for mixtures exhibiting an isotropic----smectic transition. Rotational correlation times calculated for motions about the long molecular axis and the nonunique axis showed (i) that the ring motions became more anisotropic as TLC was approached and (ii) that the motions were more anisotropic at TLC + 1 degree C for systems exhibiting a cholesteric phase than for systems exhibiting only a smectic phase. 13C line widths in spectra of the cholesteryl ester liquid-crystalline phases suggested that TO perturbed the cholesteryl ester intermolecular interactions and increased the rates of cholesteryl ester molecular motions relative to neat esters.  相似文献   

12.
The biphasic nature of the time course of the action of staphylococcal nuclease on thymus nucleohistone was confirmed by studying the hydrolysis of this nucleoprotein at various enzyme concentrations. The transition from the rapid first to the sluggish second phase of the time course was particularly distinct at the highest enzyme concentrations. The rapid initial phase of the hydrolysis curve leveled off sharply when between 60 and 65 per cent of the total TNH phosphorus had been converted to acid-soluble phosphorus compounds.The insoluble complexes of TNH with protamines were found to be very resistant against the action of staphylococcal nuclease.The time course of the action of staphylococcal nuclease on a commercial nucleoprotamine of salmon testicles was found to become very sluggish when between 35 and 40 per cent of its total phosphorus had been converted to acid-soluble phosphorus compounds.When nucleoprotamines prepared in the laboratory from the secreted sperm cell suspension of Brown Brook Trout were digested with staphylococcal nuclease, only between 15 and 20 per cent of the total phosphorus were cleaved to acid-soluble phosphorus compounds during the rapid phase of the nuclease action.The respective values for the phosphorus fractions available for magnesium-binding and those susceptible to the rapid cleavage by staphylococcal nuclease were found to be very similar.  相似文献   

13.
Isothermal-isobaric molecular dynamics simulations are used to calculate the specific volume of models of trehalose and three amorphous trehalose-water mixtures (2.9%, 4.5% and 5.3% (w/w) water, respectively) as a function of temperature. Plots of specific volume versus temperature exhibit a characteristic change in slope when the amorphous systems change from the glassy to the rubbery state and the intersection of the two regression lines provides an estimate of the glass transition temperature T(g). A comparison of the calculated and experimental T(g) values, as obtained from differential scanning calorimetry, shows that despite the predicted values being systematically higher (about 21-26K), the trend and the incremental differences between the T(g) values have been computed correctly: T(g)(5.3%(w/w))相似文献   

14.
Unlike the static length-tension curve of striated muscle, airway and urinary bladder smooth muscles display a dynamic length-tension curve. Much less is known about the plasticity of the length-tension curve of vascular smooth muscle. The present study demonstrates that there were significant increases of ~15% in the phasic phase and ~10% in the tonic phase of a third KCl-induced contraction of a rabbit femoral artery ring relative to the first contraction after a 20% decrease in length from an optimal muscle length (L(0)) to 0.8-fold L(0). Typically, three repeated contractions were necessary for full length adaptation to occur. The tonic phase of a third KCl-induced contraction was increased by ~50% after the release of tissues from 1.25-fold to 0.75-fold L(o). The mechanism for this phenomenon did not appear to lie in thick filament regulation because there was no increase in myosin light chain (MLC) phosphorylation to support the increase in tension nor was length adaptation abolished when Ca(2+) entry was limited by nifedipine and when Rho kinase (ROCK) was blocked by H-1152. However, length adaptation of both the phasic and tonic phases was abolished when actin polymerization was inhibited through blockade of the plus end of actin by cytochalasin-D. Interestingly, inhibition of actin polymerization when G-actin monomers were sequestered by latrunculin-B increased the phasic phase and had no effect on the tonic phase of contraction during length adaptation. These data suggest that for a given level of cytosolic free Ca(2+), active tension in the femoral artery can be sensitized not only by regulation of MLC phosphatase via ROCK and protein kinase C, as has been reported by others, but also by a nonmyosin regulatory mechanism involving actin polymerization. Dysregulation of this form of active tension modulation may provide insight into alterations of large artery stiffness in hypertension.  相似文献   

15.
通过计算机对心肌振荡模型的研究,得到它在单刺激下的相角转移曲线.用相角转移曲线研究了振子在不同周期刺激下表现出来的锁相、分叉及混沌现象,进而对某些常见的心律失常产生的机制进行了讨论.  相似文献   

16.
Recombinant Lactobacillus leichmannii ribonucleosidetriphosphate reductase (RTPR, E.C.1.17.4.2) constitutively expressed by E. coli HB101 pSQUIRE has been purified from sonicated cell material in a one-step procedure by PEG 4000 (16% (w/w))/phosphate (7% (w/w)) liquid-liquid extraction. A high yield of 75.1% RTPR in the top phase and a partitioning of 8.5:1 between total RTPR activity in top and bottom phase were obtained in this preparative system. The RTPR-containing top phase was used to reduce ATP in the 2'-position on a gram scale with high final conversion and yield proving the ribonucleotide reductase approach feasible for the preparative synthesis of 2'-deoxyribonucleotides. High concentrations of sodium acetate in the reaction served to substitute for allosteric effectors of RTPR. 1,4-Dithio-DL-threitol was used as an artificial reducing agent for RTPR.  相似文献   

17.
J W Fox  D P Owens  K P Wong 《Biochemistry》1978,17(8):1357-1364
The denaturation of ribosome and RNA by ethylene glycol (EG) has been studied in an attempt to further understand the conformation and stability of the ribosome. At high concentrations of EG, the ribosome, its subunits, and 16S RNA undergo drastic structural changes as shown by circular dichroism, ultraviolet absorption spectroscopy, and sedimentation velocity. Two separate conformational transitions were observed for the 30S subunit; one from 30 to 50% EG and another from 60 to 90% EG. This observation suggests the presence of two "domains" in the 30S subunit which differ in their stability. However, the 50S subunit undergoes a single sharp transition at 60 to 90% EG, consistent with the notion of a highly cooperative conformation. Association of the subunits stablizes part of the 30S subunit since the transition curve for the 70S ribosome does not exhibit significant change at the low EG concentration region as seen for the 30S subunit. Removal of proteins from the 30S subunit broadens the transition curve to lower EG concentrations and suggests the role of proteins in stabilizing the conformation of the 16S RNA.  相似文献   

18.
The effect of 100 atm pressure on the organization of the lipid-peptide complex formed between polymyxin and dipalmitoyl phosphatidic acid has been investigated. Phase transition curves were obtained by electron paramagnetic resonance by measuring the partition coefficient of the spin label, 2, 2, 5, 5-tetramethylpiperidine-N-oxyl. The three-step phase transition curve previously obtained with fluorescence polarization measurements was confirmed, demonstrating three distinct phosphatidic acid domains in the bilayer. Pressure increases binding of polymyxin to phosphatidic acid bilayers and alters the proportions of the two domains that differ in the mode of binding between phosphatidic acid and polymyxin. The binding curves of polymyxin to phosphatidic acid bilayers wre determined and it was shown that application of pressure reduces the cooperativity of the binding curve.  相似文献   

19.
Low shear viscosities have been determined for a 1 mg/ml poly(L -lysine) solution as a function of added salt concentration in the region of the previously reported ordinary–extraordinary phase transition. The measured viscosities indicate that the polyions are far from completely extended at the transition. Estimates of the longest internal relaxation time for an equivalent free-draining Rouse-Zimm chain give τ ? 10?5 sec, similar to that of the rapid, angle-independent component previously observed in the dynamic light-scattering correlation function at the transition. An unusual peak and valley are observed in the curve of [η]0 versus [NaBr] in the transition region. Possible interpretations of these features, and their bearing on the nature of the extraordinary phase, are discussed.  相似文献   

20.
Temperature cycling across the glass transition of the aqueous phase of oil-in-water emulsions stabilized by whey protein isolate was considered as a possible factor affecting stability. Emulsions were formulated with an aqueous phase containing 80% (w/w) fructose, fructose:glucose 1:1 or glucose, in order to prepare a glass forming aqueous phase with sugar concentration corresponding to that of the unfrozen phase of the maximally freeze-concentrated solutions. This allowed thermal cycling across the glass transition in the absence of the formation of ice crystals. Emulsion stability was studied using differential scanning calorimetry, dynamic light scattering and by visual analysis of the morphology of the systems. Emulsified systems undergoing glass transition cycles of the aqueous phase did not show destabilization of the dispersed (crystallized) lipid phase. Sugar crystallization in the aqueous phase, which occurred when glucose systems were stored above the Tg, led to emulsion breakdown. In this study, the formation of a glassy structure in the continuous aqueous phase preserved the interfacial structure of WPI, thus protecting the dispersed lipid phase from destabilization. On the contrary, glucose crystallization caused disruption of the interfacial membrane structure and loss of integrity of the interface which resulted in extensive lipid phase destabilization.  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