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 melting curves of 11 vegetable oils have been characterised. Vegetable oil samples that were cooled at a constant rate (5 degrees C/min) from the melt showed between one and seven melting endotherms upon heating at four different heating rates (1, 5, 10 and 20 degrees C/min) in a differential scanning calorimeter (DSC). Triacylglycerol (TAG) profiles and iodine value analyses were used to complement the DSC data. Generally, the melting transition temperature shifted to higher values with increased rates of heating. The breadth of the melting endotherm and the area under the melting peak also increased with increasing heating rate. Although the number of endothermic peaks was dependent on heating rate, the melting curves of the oil samples were not straightforward in that there was no correlation between the number of endothermic peaks and heating rates. Multiple melting behaviour in DSC experiments with different heating rates could be explained by: (1) the melting of TAG populations with different melting points; and (2) TAG crystal reorganisation effects. On the basis of the corollary results obtained, vegetable oils and fats may be distinguished from their offset-temperature (Toff) values in the DSC melting curves. The results showed that Toff values of all oil samples were significantly (p < 0.01) different in the melting curves scanned at four different scanning rates. These calorimetric results indicate that DSC is a valuable technique for studying vegetable oils.  相似文献   

2.
To survive freezing, cells must not undergo internal ice formation during cooling. One vital factor is the cooling rate. The faster cells are cooled, the more their contents supercool, and at some subzero temperature that supercooled cytoplasm will freeze. The question is at what temperature? The relation between cooling rate and cell supercooling can be computed. Two important parameters are the water permeability (Lp) and its temperature dependence. To avoid intracellular ice formation (IIF), the supercooling must be eliminated by dehydration before the cell cools to its ice nucleation temperature. With an observed nucleation temperature of −25 °C, the modeling predicts that IIF should not occur in yeast cooled at <20 °C/min and it should occur with near certainty in cells cooled at ?30 °C/min. Experiments with differential scanning calorimetry (DSC) confirmed these predictions closely. The premise with the DSC is that if there is no IIF, one should see only a single exotherm representing the freezing of the external water. If IIF occurs, one should see a second, lower temperature exotherm. A further test of whether this second exotherm is IIF is whether it disappears on repeated freezing. IIF disrupts the plasma membrane; consequently, in a subsequent freeze cycle, the cell can no longer supercool and will not exhibit a second exotherm. This proved to be the case at cooling rates >20 °C/min.  相似文献   

3.
The exothermic and endothermic peaks in cooling and heating curves of differential scanning calorimetry (DSC) for gellan gum gels without and with potassium chloride and sodium chloride were analyzed. The gelling and melting temperatures shifted to higher temperatures with increasing gellan and salt concentration in the concentration range of gellan from 0.3 to 2.0% (w/w). The exothermic and endothermic enthalpy increased with increasing gellan and salt concentrations. Cooling DSC curves showed one exothermic peak for samples with salts and at low gellan concentration. Heating DSC curves showed many peaks for all samples except 0.3% (w/w) gellan gum gels. The sol-gel transition of samples was examined numerically by using a zipper model approach. The introduction of cations increases the number of junction zones or zippers and decreases the rotational freedom of parallel links. This makes the structure of junction zones more heat resistant, and increases the elastic modulus of the gel.  相似文献   

4.
The purpose of this work was to determine what aspect of the milled compound influences its thermal profile. For this, six different compounds with different properties were chosen and cryomilled for different times to get an amorphous solid. Differential scanning calorimetry (DSC) and X-ray powder diffraction were used to characterize the material and look at the thermal behavior. Melt-quenched samples were also prepared, and the thermal profile upon milling was determined and correlated with the thermal behavior of the cryomilled samples. Growth rates were determined by hot-stage microscopy. Ketoconazole, when cryomilled, showed only one crystallization exotherm in the DSC profile. Ursodiol, and to some extent indomethacin, initially showed a double exotherm which eventually become a single exotherm on further milling. Griseofulvin, carbamazepine, and piroxicam exhibited a double exotherm in the DSC profile upon cryomilling to the amorphous state. Surface crystal growth rates around T (g) were found to be highest for compounds showing the double exotherm in the DSC. Thus, it was seen that compounds which have high surface crystallization tendency will exhibit the double exotherm during heating.  相似文献   

5.
The freeze-preservation of pollen is dependent on the interaction of several factors such as freezing rate, thawing rate, freeze-drying temperature and duration, storage temperature and environment and rehydration rates. Changes in any of these variables affects the others directly or indirectly.Rapid freezing of pollen at rates of approximately 200 °C/min maintains the highest degree of viable pollen in combination with rapid thawing rates of 218 °C/min. Rapid cooling and slow rewarming resulted in a substantial loss of pollen viability. This might indicate that intracellular ice crystals formed during rapid cooling perhaps grow into larger ice masses during slow rewarming or storage at temperatures above ?50 °C.The germinability of pollen freeze-dried at temperatures below ?50 °C was also prolonged over that of the controls. Germination values for unfrozen pollen stored for 30 days at 0–5 °C averaged 50% for lily and 20% for corn. Freeze-dried pollen stored for 30 days at the same temperature yielded considerably higher viability percentages for both lily and corn pollen. Drying time is an important factor, perhaps indicating that residual moisture is critical. Freeze-dried pollen can be stored at higher temperatures than frozen and control pollen. Freeze-dried material stored for five months at 0–5 °C, upon slow rehydration yielded intact grains which has average germination percentages of 25 for lily and 15 for corn. The same pollen upon rapid rehydration showed rupturing of 20–40% of the cells and practically no germination.  相似文献   

6.
Solutions of antifreeze glycoproteins 1 through 5 and 8 were analyzed for activity by differential scanning calorimetry. With a scan rate of 1 degree C min-1, antifreeze glycoproteins 1-5 (20 mg/ml) revealed antifreeze activity with a delay in the freeze exotherm during cooling in the presence of ice. Antifreeze glycoprotein 8 (60 mg/ml), however, did not reveal antifreeze activity. When a 0.1 degree C min-1 scan rate was used, glycoproteins 1-5 again yielded a delay in the freeze onset, but the exotherm consisted of multiple events. At the slower scan glycoprotein 8 revealed an initial freeze followed by multiple exothermic events resembling those of glycoproteins 1-5. Thermograms exhibiting antifreeze activity had an initial shoulder in the exotherm direction upon cooling followed by a delay before the exotherm. The shoulders were correlated with c-axis ice growth observed in visual methods. The glycoprotein antifreezes had a linear increase in activity with decreased ice content.  相似文献   

7.
S Matuoka  H Yao  S Kato    I Hatta 《Biophysical journal》1993,64(5):1456-1460
In the ripple phase of fully hydrated multilamellar vesicles of dipalmitoylphosphatidylcholine (DPPC), two kinds of small-angle x-ray diffraction profiles are observed on cooling through the main transition. One is a seemingly normal profile similar to that observed on heating and the other is the superposition of the diffraction profiles for the primary (normal) and the secondary ripple structures. We found that the profile obtained depended on the cooling rate. Increasing the cooling rate from 0.1 degrees C/min to 1 degrees C/min caused the peaks originating from the secondary ripple structure to diminish. After a cooling scan at 43 degrees C/min, the profile became similar to that of the normal ripple structure, although a trace of the secondary ripple structure remains. The results are interpreted in terms of the rise and fall of three-dimensional correlated domains composed of both primary and secondary ripple structures. At slow cooling rates, correlated domains of both kinds of ripple structures develop. As the cooling rate is increased, the domain of the primary ripple structure remains correlated, while that of the secondary ripple structure becomes less correlated. In addition, the multipeak profile appears even at rapid cooling rates, if the final low temperature lies just below the Tm for the main transition. This results suggests that formation of the correlated domains of the secondary ripple structure requires a certain time interval during which the DPPC vesicles experience the temperature just below the main transition. The secondary ripple structure takes place in phosphatidylcholines having more than 15 carbons in each hydrocarbon chain upon cooling through the main transition.  相似文献   

8.
A three-part, coupled model of cell dehydration, nucleation, and crystal growth was used to study intracellular ice formation (IIF) in cultured hepatocytes frozen in the presence of dimethyl sulfoxide (DMSO). Heterogeneous nucleation temperatures were predicted as a function of DMSO concentration and were in good agreement with experimental data. Simulated freezing protocols correctly predicted and explained experimentally observed effects of cooling rate, warming rate, and storage temperature on hepatocyte function. For cells cooled to -40 degrees C, no IIF occurred for cooling rates less than 10 degrees C/min. IIF did occur at faster cooling rates, and the predicted volume of intracellular ice increased with increasing cooling rate. Cells cooled at 5 degrees C/min to -80 degrees C were shown to undergo nucleation at -46.8 degrees C, with the consequence that storage temperatures above this value resulted in high viability independent of warming rate, whereas colder storage temperatures resulted in cell injury for slow warming rates. Cell damage correlated positively with predicted intracellular ice volume, and an upper limit for the critical ice content was estimated to be 3.7% of the isotonic water content. The power of the model was limited by difficulties in estimating the cytosol viscosity and membrane permeability as functions of DMSO concentration at low temperatures.  相似文献   

9.
The objective of this study was to optimize interrupted slow-freezing protocols for African catfish semen. Semen diluted with methanol and extender was frozen in 1-ml vials in a programmable freezer. The temperatures of the freezer (T(chamber)) and of the semen (T(semen)) were measured simultaneously. We first tested two-step freezing protocols with different cooling rates (-2, -5, and -10 degrees C/min) and different temperatures at plunging into liquid N2. The difference between T(semen) and T(chamber) increased with faster cooling rates. In all programs, survival of spermatozoa, expressed as hatching rates, increased from near zero when T(semen) at plunging was higher than -30 degrees C to values equal to those of control when T(semen) at plunging was equal to or lower than -38 degrees C. The inclusion of an isothermal holding period before plunging into liquid N2 (three-step freezing protocols) resulted in an equilibration between T(semen) and T(chamber) and improved semen survival. Semen could be plunged at temperatures as high as -36 degrees C when cooled at -5 or -10 degrees C/min, without compromising postthaw semen survival. Cooling at -2 degrees C/min in combination with a 5-min holding period reduced postthaw survival. We conclude that with slow cooling rates of -2 to -5 degrees C/min, hatching rates can be maximized by plunging as soon as T(semen) reaches -38 degrees C. The isothermal holding period is beneficial when faster rates are used. A simple and efficient protocol for freezing African catfish semen can be obtained by cooling at a rate of -5 to -10 degrees C/min combined with a 5-min holding period in the freezer, at -40 degrees C.  相似文献   

10.
Babesia rodhaini parasites in murine blood containing 1.5 m DMSO were frozen at two rates, as judged by the duration of the “freezing plateau”, then cooled to ?196 °C and rewarmed at two rates to detect interactions between the duration of the plateau and rates of subsequent cooling and rewarming. Infectivity tests showed that fast and slow freezing (plateau times of about 1 sec and 30 sec, respectively) had similar effects on parasite survival when cooling was at 130 °C/min and warming was at 800 °C/min. However, when either the cooling rate was increased to 3500 °C/min or the warming rate was decreased to 2.3 °C/min, fast freezing decreased parasite survival more than did slow freezing. It is suggested that fast freezing accentuated the damaging effects of fast cooling and slow warming by increasing intracellular ice formation.  相似文献   

11.
Cryomicroscopy and differential scanning calorimetry (DSC) were used to characterize the incidence of intracellular ice formation (IIF) in 12- to 13-hr-old embryos of Drosophila melanogaster (Oregon-R strain P2) as influenced by the state of the eggcase (untreated, dechorionated, or permeabilized), the composition of the suspending medium (with and without cryoprotectants), and the cooling rate. Untreated eggs underwent IIF over a very narrow temperature range when cooled at 4 or 16 degrees C/min with a median temperature of intracellular ice formation (TIIF50) of -28 degrees C. The freezable water volume of untreated eggs was approximately 5.4 nl as determined by DSC. IIF in dechorionated eggs occurred over a much broader temperature range (-13 to -31 degrees C), but the incidence of IIF increased sharply below -24 degrees C, and the cumulative incidence of IIF at -24 degrees C decreased with cooling rate. In permeabilized eggs without cryoprotectants (CPAs), IIF occurred at much warmer temperatures and over a much wider temperature range than in untreated eggs, and the TIIF50 was cooling rate dependent. At low cooling rates (1 to 2 degrees C/min), TIIF50 increased with cooling rate; at intermediate cooling rates (2 to 16 degrees C/min), TIIF50 decreased with cooling rate. The total incidence of IIF in permeabilized eggs was 54% at 1 degree C/min, and volumetric contraction almost always occurred during cooling. Decreasing the cooling rate to 0.5 degree C/min reduced the incidence of IIF to 43%. At a cooling rate of 4 degrees C/min, ethylene glycol reduced the TIIF50 by about 12 degrees C for each unit increase in molarity of CPA (up to 2.0 M) in the suspending medium. The TIIF50 was cooling rate dependent when embryos were preequilibrated with 1.0 M propylene glycol or ethylene glycol, but was not so in 1.0 M DMSO. For embryos equilibrated in 1.5 M ethylene glycol and then held at -5 degrees C for 1 min before further cooling at 1 degree C/min, the incidence of IIF was decreased to 31%. Increasing the duration of the isothermal hold to 10 min reduced the incidence of IIF to 22% and reduced the volume of freezable water in embryos when intracellular ice formation occurred. If the isothermal hold temperature was -7.5 or -10 degrees C, a 10- to 30-min holding time was required to achieve a comparable reduction in the incidence of IIF.  相似文献   

12.
The current study presents a new and novel analysis of heat release signatures measured by a differential scanning calorimeter (DSC) associated with water transport (WT), intracellular ice formation (IIF) and extracellular ice formation (EIF). Correlative cryomicroscopy experiments were also performed to validate the DSC data. The DSC and cryomicroscopy experiments were performed on human dermal fibroblast cells (HDFs) at various cytocrit values (0–0.8) at various cooling rates (0.5–250 °C/min). A comparison of the cryomicroscopy experiments with the DSC analysis show reasonable agreement in the water transport (cellular dehydration) and IIF characteristics between both the techniques with the caveat that IIF measured by DSC lagged that measured by cryomicroscopy. This was ascribed to differences in the techniques (i.e. cell vs. bulk measurement) and the possibility that not all IIF is associated with visual darkening. High and low rates of 0.5 °C/min and 250 °C/min were chosen as HDFs did not exhibit significant IIF or WT at each of these extremes respectively. Analysis of post-thaw viability data suggested that 10 °C/min was the presumptive optimal cooling rate for HDFs and was independent of the cytocrit value. The ratio of measured heat values associated with IIF (qIIF) to the total heat released from both IIF and water transport or from the total cell water content in the sample (qCW) was also found to increase as the cooling rate was increased from 10 to 250 °C/min and was independent of the sample cytocrit value. Taken together, these observations suggest that the proposed analysis is capable of deconvolving water transport and IIF data from the measured DSC latent heat thermograms in cell suspensions during freezing.  相似文献   

13.
Wang HS  Kang L 《Cryobiology》2005,51(2):220-229
To examine the relationship between cooling rate and cold hardiness in eggs of the migratory locust, Locusta migratoria, the survival rates and cryoprotectant levels of three embryonic developmental stages were measured at different cooling rates (from 0.05 to 0.8 degrees C min(-1)) in acclimated and non-acclimated eggs. Egg survival rate increased with decreasing cooling rate. The concentration of cryoprotectants (myo-inositol, trehalose, mannitol, glycerol, and sorbitol) increased in non-acclimated eggs, but varied significantly in response to different cooling rates in acclimated eggs. The acclimation process (5 degrees C for 3 days) did not increase eggs resistance to quick cooling ("plunge" cooling and 0.8 degrees C min(-1)). Earlier stage embryos were much more sensitive than later stage embryos to the same cooling rates. Time spent at subzero temperatures also had a strong influence on egg survival.  相似文献   

14.
The effect of low muscle temperature on the response to dynamic exercise was studied in six healthy men who performed 42 min of exercise on a cycle ergometer at an intensity of 70% of their maximal O2 uptake. Experiments were performed under control conditions, i.e. from rest at room temperature, and following 45 min standing with legs immersed in a water bath at 12 degrees C. The water bath reduced quadriceps muscle temperature (at 3 cm depth) from 36.4 (SD 0.5) degrees C to 30.5 (SD 1.7) degrees C. Following cooling, exercise heart rate was initially lower, the mean difference ranged from 13 (SD 4) beats.min-1 after 6 min of exercise, to 4 (SD 2) beats.min-1 after 24 min of exercise. Steady-state oxygen uptake was consistently higher (0.2 l.min-1). However, no difference could be discerned in the kinetics of oxygen uptake at the onset of exercise. During exercise after cooling a significantly higher peak value was found for the blood lactate concentration compared to that under control conditions. The peak values were both reached after approximately 9 min of exercise. After 42 min of exercise the blood lactate concentrations did not differ significantly, indicating a faster rate of removal during exercise after cooling. We interpreted these observations as reflecting a relatively higher level of muscle hypoxia at the onset of exercise as a consequence of a cold-induced vasoconstriction. The elevated steady-state oxygen uptake may in part have been accounted for by the energetic costs of removal of the extra lactate released into the blood consequent upon initial tissue hypoxia.  相似文献   

15.
Seeded solutions of catalase in neutral 10 mM potassium phosphate buffer exhibited characteristic rate dependencies for freeze-thaw damage: Damage increased as the cooling rate was increased, and as the warming rate was decreased. The pattern of warming-rate dependence was independent of the prior cooling rate and also of the addition of KCl or of NaCl to the buffer. In contrast, the cooling-rate curve became almost flat upon addition of 0.1 M KCl, suggesting increased damage from concentrating solute at low cooling rates. In the presence of added NaCl, frank optimum-recovery cooling-rate curves were generated. At low NaCl levels (less than 10 mM) the optimum occurred at 0.5 °C/ min; at 27 and 81 mM NaCl, the optimum shifted to 5 and 20 °C/min, respectively. By comparison with KCl, it appears that the major factor causing damage at low cooling rates in NaCl is acidification. The factor causing damage at high cooling rates remains obscure. The argument that it is due to the trapping of the enzyme molecules at interfaces at high dilution, to be subsequently damaged by shearing stress or dehydration during the recrystallization attending slow warming, is mitigated by the finding that inactivation remains a function of the initial enzyme concentration at all cooling rates. The possibility that a particular conformational state is trapped in an unfavorable temperature zone was also considered: Three simple models were formulated, and the relative order of recovery was deduced for the possible sequences of fast and slow cooling and warming. The permutation observed for catalase was inconsistent with any of these three mechanisms, although they may be pertinent for the red cell and other systems. A final possibility, not yet explored, is that rapid cooling causes damage by producing nonequilibrium freezing, with large deviations of pH and/or solute concentration from those expected at equilibrium.  相似文献   

16.
To model the cryobiological responses of cells and tissues, permeability characteristics are often measured at suprazero temperatures and the measured values are used to predict the responses at subzero temperatures. The purpose of the present study was to determine whether the rate of cooling from +25 to +4 degrees C influenced the measured water transport response of ovarian tissue at subzero temperatures in the presence or absence of cryoprotective agents (CPAs). Sections of freshly collected equine ovarian tissue were first cooled either at 40 degrees C/min or at 0.5 degrees C/min from 25 to 4 degrees C, and then cooled to subzero temperatures. A shape-independent differential scanning calorimeter (DSC) technique was used to measure the volumetric shrinkage during freezing of equine ovarian tissue sections. After ice was induced to form in the extracellular fluid within the specimen, the sample was frozen from the phase change temperature to -50 degrees C at 5 degrees C/min. Replicate samples were frozen in isotonic medium alone or in medium containing 0.85 M glycerol or 0.85 M dimethylsulfoxide. The water transport response of ovarian tissue samples cooled at 40 degrees C/min from 25 to 4 degrees C was significantly different (confidence level >95%) from that of tissue samples cooled at 0.5 degrees C/min, whether in the presence or absence of CPAs. We fitted a model of water transport to the experimentally-derived volumetric shrinkage data and determined the best-fit membrane permeability parameters (L(pg) and E(Lp)) of equine ovarian tissue during freezing. Subzero water transport parameters of ovarian tissue samples cooled at 0.5 degrees C/min from 25 to 4 degrees C ranged from: L(pg) = 0.06 to 0.73 microm/min.atm and E(Lp) = 6.1 to 20.5 kcal/mol. The corresponding parameters of samples cooled at 40 degrees C/min from 25 to 4 degrees C ranged from: L(pg) = 0.04 to 0.61 microm/min.atm and E(Lp) = 8.2 to 54.2 kcal/mol. Calculations made of the theoretical response of tissue at subzero temperatures suggest that the optimal cooling rates to cryopreserve ovarian tissue are significantly dependent upon suprazero cooling conditions.  相似文献   

17.
Twelve saturated mixed-chain phosphatidylcholines have been identified for which the thermotropic phase behavior observed upon cooling from the L alpha phase is dependent upon the thermal history of the sample in the gel phase. If fully hydrated samples of these lipids are cooled and soon thereafter examined by differential scanning calorimetry, one observes a single highly cooperative endotherm (the chain-melting phase transition) upon heating, and on subsequent cooling, a single exotherm that may occur at temperatures as much as 4-6 degrees C below that of the single endotherm observed upon heating. In contrast, if the samples are incubated in the gel state at low temperatures for prolonged periods of time, one observes a single heating endotherm as before, but two sharp exotherms upon cooling. The latter transitions occur at temperatures close to that of the single endotherm observed upon heating and the single cooling exotherm observed prior to incubation in the gel state. The combined enthalpy of the two cooling exotherms is the same as that of the single heating endotherm or the single cooling exotherm initially observed. Infrared spectroscopic and X-ray diffraction studies indicate that the structural conversions characteristic of liquid-crystalline/gel phase transitions occur at both of those cooling exotherms. Of the 12 lipids that exhibit this unusual behavior, nine fulfill the previously defined structural requirements for the formation of the so-called mixed-interdigitated gel phase, and there is evidence in the literature that one of the three remaining lipids also forms such a structure. Infrared spectroscopic studies of the other two lipids indicate that their gel phases exhibit spectroscopic features that closely resemble those of lipids that meet the previously defined structural criteria for the formation of mixed-interdigitated gel phases and that differ markedly from those of both saturated symmetric-chain and saturated mixed-chain phosphatidylcholines that do not normally form mixed-interdigitated gel phases. Also, electron density reconstructions based on small-angle X-ray diffraction studies of the gel phases of those two lipids indicate that the thickness of their gel phase bilayers is consistent with their forming mixed-interdigitated gel phases. Thus the unusual thermotropic phase behavior described here may be a general characteristic of phosphatidylcholines that form mixed-interdigitated gel phases. This unusual behavior is not associated with any major change in any of several physical properties of these lipid bilayers but may arise from an alteration of the size and/or structure of microdomains present in the liquid-crystalline phase.  相似文献   

18.
于瑞凤  朱建军 《植物学报》2018,53(2):203-211
低温条件下植物组织的结冰模式, 即胞外或胞内结冰, 直接决定着细胞的生死。目前胞外结冰的直接证据很少, 尤其缺乏热力学证据。用高分辨率差热扫描结合显微观察分析了女贞(Ligustrum lucidum)和冬青卫矛(Euonymus japonicus)叶片在降温过程中的结冰放热现象, 发现2种植物胞外结冰的热力学和组织结构变化的新证据。2种植物的活叶片在冷却过程中均呈现双放热峰, 即双相结冰的特征; 而预先冰冻杀死的叶片和叶片组织提取液浸润的滤纸片在同样冷却过程中仅有1个快速的大单放热峰, 即单相结冰的特征。显微观察也显示, 结冰过程中活组织细胞间隙中形成大量的白色冰晶, 且细胞虽然脱水收缩但细胞内的有色溶液没有流失, 表现出胞外结冰的特征。实验结果为深入揭示植物的冰冻伤害机制提供了新证据和研究方法。  相似文献   

19.
A tissue culture assay has been used to measure the survival of murine lymphoma cells (L-cells) after freezing and thawing in the presence of 2 M glycerol or 1.6 M dimethyl sulfoxide. The effect of variations in cooling rate (0.1 to 10.0 °C/min) and warming rate (0.3 to 200 °C/min) were studied. It was found that survival exhibited a peak at the “conventional” combination of slow cooling and rapid warming (~1 and 200 °C/ min, respectively). It was also shown, however, that a second peak of similar magnitude occurred when the cells were cooled and rewarmed at 0.2-0.3 °C/min. These results are interpreted on the basis of current theories of freezing injury, stressing the importance of damage produced by the recrystallization of intracellular ice and by solute loading. The ultraslow rates of cooling and rewarming which produced the second survival peak are practicable for whole organs, and their potential importance for organ cryopreservation is apparent.  相似文献   

20.
The binary phase behavior of pure 1,3-dimyristoyl-2-stearoyl-sn-glycerol (MSM) and 1,2-dimyristoyl-3-stearoyl-sn-glycerol (MMS) was investigated in terms of polymorphism, melting and crystallization behavior, SFC, hardness and microstructure. Samples were crystallized at cooling rates of 3.0 and 0.1 degrees C/min. The asymmetric TAG demonstrated lower melting and crystallization points at both cooling rates. All samples crystallized in the beta' polymorph when cooled at 0.1 degrees C/min and in the alpha polymorph when cooled at 3.0 degrees C/min. The experimentally determined kinetic phase diagram of MSM-MMS was monotectic for both cooling rates. This data was well described by a thermodynamic model using the Bragg-Williams approximation for non-ideality of mixing and suggested that in both the solid and liquid states, like pair interactions (MSM-MSM and MMS-MMS) were favored over MSM-MMS interaction. A strong tendency to phase separation in the solid phase was also observed. For both cooling rates, the fit of the SFC (%)-time curves to a modified form of the Avrami model indicated that crystallization occurred in two distinct kinetic steps. Depressions seen in SFC did not correspond to depressions in hardness or melting temperatures.  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