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.
Ice formation on warming is of comparable or greater importance to ice formation on cooling in determining survival of cryopreserved samples. Critical warming rates required for ice-free warming of vitrified aqueous solutions of glycerol, dimethyl sulfoxide, ethylene glycol, polyethylene glycol 200 and sucrose have been measured for warming rates of order 10–104 K/s. Critical warming rates are typically one to three orders of magnitude larger than critical cooling rates. Warming rates vary strongly with cooling rates, perhaps due to the presence of small ice fractions in nominally vitrified samples. Critical warming and cooling rate data spanning orders of magnitude in rates provide rigorous tests of ice nucleation and growth models and their assumed input parameters. Current models with current best estimates for input parameters provide a reasonable account of critical warming rates for glycerol solutions at high concentrations/low rates, but overestimate both critical warming and cooling rates by orders of magnitude at lower concentrations and larger rates. In vitrification protocols, minimizing concentrations of potentially damaging cryoprotectants while minimizing ice formation will require ultrafast warming rates, as well as fast cooling rates to minimize the required warming rates.  相似文献   

2.
Thermal properties of ethylene glycol aqueous solutions   总被引:3,自引:0,他引:3  
Baudot A  Odagescu V 《Cryobiology》2004,48(3):283-294
Preventing ice crystallization by transforming liquids into an amorphous state, vitrification can be considered as the most suitable technique allowing complex tissues, and organs cryopreservation. This process requires the use of rapid cooling rates in the presence of cryoprotective solutions highly concentrated in antifreeze compounds, such as polyalcohols. Many of them have already been intensively studied. Their glass forming tendency and the stability of their amorphous state would make vitrification a reality if their biological toxicity did not reduce their usable concentrations often below the concentrations necessary to vitrify organs under achievable thermal conditions. Fortunately, it has been shown that mixtures of cryoprotectants tend to reduce the global toxicity of cryoprotective solutions and various efficient combinations have been proposed containing ethanediol. This work reports on the thermal properties of aqueous solutions with 40, 43, 45, 48, and 50% (w/w) of this compound measured by differential scanning calorimetry. The glass forming tendency and the stability of the amorphous state are evaluated as a function of concentration. They are given by the critical cooling rates v(ccr)above which ice crystallization is avoided, and the critical warming rates v(cwr) necessary to prevent ice crystallization in the supercooled liquid state during rewarming. Those critical rates are calculated using the same semi-empirical model as previously. This work shows a strong decrease of averaged critical cooling and warming rates when ethanediol concentration increases, V(ccr) and V(cwr) = 1.08 x 10 (10) K/min for 40% (w/w) whereas V(ccr) = 11 and V(cwr) = 853 K/min for 50% (w/w). Those results are compared with the corresponding properties of other dialcohols obtained by the same method. Ethylene glycol efficiency is between those of 1,2-propanediol and 1,3-propanediol.  相似文献   

3.
The influence of hydroxyethyl starch on ice formation in aqueous solutions   总被引:2,自引:0,他引:2  
Differential scanning calorimetry, and, in some supplementary experiments, X-ray diffractometry and cryomicroscopy, were applied to study the influence of concentration (< 70 wt%) and cooling/warming rates (< 320 K/min) on ice formation in aqueous solutions of HES. The calorimetric measurements of the quantity of crystallizing water indicated that a mass fraction ? = 0.522 (i.e., grams water per gram HES) remained unfrozen. These results are in good agreement with our earlier extrapolations from ternary phase diagram data and tend to support the proposed cryoprotective mechanism. The value of ? determined during warming was essentially independent of composition up to the corresponding saturation concentration. It was observed that solutions containing 60 wt% HES or more remained wholly amorphous during cooling even at rates as low as 2.5 K/min (down to 120 K). Such glassy solutions are subject to devitrification at temperatures Td which depend on the warming rate. The concentrations close to 55 wt% HES mark a transitional range exhibiting two crystallization peaks, probably due to different mechanisms of nucleation, the portion of ice formed during cooling being related to the imposed cooling rate. All samples showed a recrystallization transition at 257.5 K which was also observed cryomicroscopically. Glass transitions, however, could not be detected by the methods applied in this study. The X-ray diffraction patterns contained the structure of only one solid phase, namely hexagonal ice. A comparison of various modifications of HES, PEG, and PVP involving bound water and melting temperature did not reveal marked differences. Minimum initial HES concentrations preventing lethal salt enrichment were computed for both binary and ternary mass fractions of NaCl as biologically relevant parameters, yielding 24.1 and 10.8 wt% HES, respectively.  相似文献   

4.
Hybridization solutions containing chaotropes may be used to modulate the thermal stability (Tm or Td) of oligodeoxynucleotide (ODN) duplexes or hybrids over a 90 degrees C range. Modulation of Td allows formulation of hybridization solutions that permit ambient temperature hybridization using most combinations of probe length, probe composition, target type, and facilitates development of convenient and rapid assay formats. The conditions required to achieve ODN duplex fidelity, and optimal yields of hybridized product, are described for trichloroacetate, thiocyanate, guanidinium salts and other chaotropic salts. The effects of different solid supports on Td are described. Also, a method is presented that uses chaotropic compounds to reduce background arising from signal ODN probes in a sandwich assay hybridization format.  相似文献   

5.
Electrostatic forces at helix-coil boundaries in DNA   总被引:2,自引:0,他引:2  
The Tm of internal loop-forming (dA.dT)N domains in pBR322 DNA has been measured over a tenfold range of [Na+]. The slopes SN = dTm/d log [Na+] are linear and decrease in magnitude with decreasing loop size N, signaling a reduction in Na+ released during the transition of these domains to the coil state. Values of SN decrease linearly with increasing N-1 in accordance with the expectation of a simple model for the occurrence of a gradient of long-range electrostatic forces at helix-coil boundaries, and extrapolate almost precisely to the value of S infinity observed for (dA.dT) infinity. These results indicate (1) less counterion is released per phosphate residue from the finite loop than from the infinite-sized loop, and (2) the difference in binding is constant for each boundary formed and independent of the size of the loop within the range examined: approximately 350 base pair (bp) greater than N greater than 71 bp. The slope of the dependence of SN on N-1 indicates the region of higher charge density at the boundary extends at least 18 A into the coil and probably 40-50 A before dropping to a value characteristic of the unperturbed coil. The free energy for excess counterion binding at boundaries can be expressed by -delta G/RT = 10.47 log[Na+] + 5.234 When the loop entropy function in a statistical mechanical algorithm for the dissociation of DNA is weighted by this quantity, calculated Tm are seen to vary by only +/- 0.09 degrees C from observed.  相似文献   

6.
Chinese Hamster Ovary (CHO) cells were cooled at 1 and 200 °C/min and subsequently thawed, while being studied with a cryomicroscope. Post-thaw size changes were measured with a Quantimet 720 Image Analysing Computer. It was found that the behavior of individuals in a population varied and depended on cooling rate. Cooling at 1 °C/min resulted in cells showing no intracellular ice, whereas cooling at 200 °C/min caused intracellular ice formation in some cells but not in others. In addition, at the slow rate, during cooling, the cells shrank significantly but swelled on thawing to become larger than non-frozen controls. Following swelling, as their temperature rose, the cells shrank to the size of non-frozen controls. At the fast rate, cells showed variation in their amount of intracellular ice and in their degree of shrinkage. Cells containing most ice shrank least. On warming, cells with intracellular ice began to swell at a lower temperature than did those cells without intracellular ice, while after thawing they swelled to a greater extent partly due to widespread blebbing. Corresponding recovery indices were measured, and correlation of these with the above effects suggests that: (i) cells completely filled with intracellular ice are non-viable; (ii) cells partially filled with intracellular ice respond to, or can be rescued by, first warming; (iii) cells without intracellular ice are viable; (iv) viable cells are those which regain their original size following thawing; (v) non-viable cells are those which remain swollen above their original size.  相似文献   

7.
Preventing intracellular ice formation is essential to cryopreserve cells. Prevention can be achieved by converting cell water into a non-crystalline glass, that is, to vitrify. The prevailing belief is that to achieve vitrification, cells must be suspended in a solution containing a high concentration of glass-inducing solutes and cooled rapidly. In this study, we vitrified 1-cell mouse embryos and examined the effect of the cooling rate, the warming rate, and the concentration of cryoprotectant on cell survival. Embryos were vitrified in cryotubes. The vitrification solutions used were EFS20, EFS30, and EFS40, which contained ethylene glycol (20, 30 and 40% v/v, respectively), Ficoll (24%, 21%, and 18% w/v, respectively) and sucrose (0.4 0.35, and 0.3 M, respectively). A 5-μl EFS solution suspended with 1-cell embryos was placed in a cryotube. After 2 min in an EFS solution at 23 °C, embryos were vitrified by direct immersion into liquid nitrogen. The sample was warmed at 34 °C/min, 4,600 °C/min and 6,600 °C/min. With EFS40, the survival was low regardless of the warming rate. With EFS30 and EFS20, survival was also low when the warming rate was low, but increased with higher warming rates, likely due to prevention of intracellular ice formation. When 1-cell embryos were vitrified with EFS20 and warmed rapidly, almost all of the embryos developed to blastocysts in vitro. Moreover, when vitrified 1-cell embryos were transferred to recipients at the 2-cell stage, 43% of them developed to term. In conclusion, we developed a vitrification method for 1-cell mouse embryos by rapid warming using cryotubes.  相似文献   

8.
All the aqueous solutions of linear saturated polyalcohols with four carbons have been investigated at low temperature. Only ice has been observed in the solutions of 1,3-butanediol and 1,2,3- and 1,2,4-butanetriol. For same solute concentration, the glass-forming tendency on cooling is highest with 2,3-butanediol, where it is comparable to that with 1,2-propanediol, the best solute reported to date. However, the quantity of ice and hydrate crystallized is particularly high on slow cooling or on subsequent rewarming. The highest stability of the amorphous state is observed on rewarming the 1,2-butanediol and 1,3-butanediol solutions. With respect to this property, these compounds come just after 1,2-propanediol and before all the other compounds studied so far. They are followed by dimethylsulfoxide and 1,2,3-butanetriol. The glass-forming tendency of the 1,3-butanediol solutions is also very high; it is third only to that of 1,2-propanediol and 2,3-butanediol. The glass-forming tendency is a little smaller with 1,2-butanediol, but it is cubic instead of ordinary hexagonal ice which crystallizes on cooling rapidly with 35% 1,2-butanediol. Cubic ice is thought to be innocuous. A gigantic glass transition is observed with 45% of this strange solute. 1,4-Butanediol, 45% also favors cubic ice greatly. Therefore, 1,2- and 1,3-butanediol with comparable physical properties are perhaps as interesting as 1,2-propanediol for cryopreservation of cells or organs by complete vitrification. Together with 1,2-propanediol, 1,2- and 1,3-butanetriol, 1,2,3-butanetriol, and perhaps 2,3-butanediol provide an interesting battery of solutions for cryopreservation by vitrification.  相似文献   

9.
X-ray structure analysis, IR spectrometry, differential thermal analysis, and viscosimetry have been used to investigate the properties of novel five-component polyhydroxyalkanoates formed by short- and medium-chain-length monomers synthesized by the bacterium Wautersia eutropha B5786. As the molar fraction of hydroxyhexanoate contained in polyhydroxyalkanoates samples increased from 2.5 to 18.0 mol%, their degree of crystallinity decreased from 72 to 57%. The melting temperature of multicomponent polyhydroxyalkanoates (Tm) and their temperature for the onset of decomposition (Td) are lower than those of polyhydroxybutyrate, whose Tm is 168-170 degrees C and Td 260-265 degrees C. In multicomponent polymers (PHA(SC+MC)), both parameters decrease as the molar fraction of hydroxyhexanoate grows to 156 and 252 degrees C, respectively, in the range of hydroxyhexanoate content studied. Hydroxyhexanoate influences the physicochemical properties of polyhydroxyalkanoates similarly to hydroxyvalerate; as the fraction of either of these medium-chain-length monomers in polyhydroxyalkanoates increases, the crystallinity of the polymer decreases, but its thermostability remains unchanged.  相似文献   

10.
Populations vary in time and in space, and temporal variation may differ from spatial variation. Yet, in the past half century, field data have confirmed both the temporal and spatial forms of Taylor's power Law, a linear relationship between log(variance) and log(mean) of population size. Recent theory predicted that competitive species interactions should reduce the slope of the temporal version of Taylor's Law. We tested whether this prediction applied to the spatial version of Taylor's Law using simple, well-controlled laboratory populations of two species of bacteria that were cultured either separately or together for 24 h in media of widely varying nutrient richness. Experimentally, the spatial form of Taylor's Law with a slope of 2 held for these simple bacterial communities, but competitive interactions between the two species did not reduce the spatial Taylor's Law slope. These results contribute to the widespread usefulness of Taylor's Law in population ecology, epidemiology and pest control.  相似文献   

11.
12.
Cryoprotection of red blood cells by 1,3-butanediol and 2,3-butanediol   总被引:1,自引:1,他引:0  
P Mehl  P Boutron 《Cryobiology》1988,25(1):44-54
1,3-Butanediol and 2,3-butanediol have been used in buffered solutions with 20, 30, or 35% (w/w) alcohol to cool erythrocytes to -196 degrees C at different cooling rates between 1 to 3500 degrees C/min, followed by slow or rapid rewarming. 1,3-butanediol shows the same shapes of red blood cell survival curves as 1,2-propanediol. Having nearly the same physical properties, they have comparable effects on cell survival. The classical maximum of survival for intermediate cooling rates and an increase for the highest cooling rates are observed. This increase seems to be correlated with the glass-forming tendency of the solution. After the fastest cooling rates, a warming rate of 5000 degrees C/min is sufficient to avoid cell damage, but a warming rate of 100-200 degrees C/min is not. Yet both of these rates would be insufficient to avoid the intracellular ice crystallization on warming. The damage on warming after fast cooling seems once again to be correlated with the transition from cubic to hexagonal ice. For all our results, 1,3-butanediol is like a "second" 1,2-propanediol and could be useful as a cryoprotectant for preservation by total vitrification. 2,3-Butanediol always gives extremely low survival rates, though it presents good physical properties. The crystallization of its hydrate seems to be lethal on cooling or on rewarming.  相似文献   

13.
In vitrified solutions, ice can form during warming if the concentration of the cryoprotectant is insufficient. For the cryopreservation of cells, ice is innocuous when it remains outside the cell, but intracellular ice (ICI) is lethal. We tried to estimate the conditions in which ICI forms in vitrified mouse morulae during warming. The solutions for the experiments (EFS10–EFS50) contained 10–50% ethylene glycol plus Ficoll plus sucrose. When vitrified EFS20, EFS30, and EFS40 were kept at −80 °C, they remained transparent after 3 min, but turned opaque after 60 min (EFS20, EFS30) or 24 h (EFS40). Morulae were vitrified with EFS solutions after exposure for 30–120 s at 25 °C. They were warmed by various methods and survival was assessed in culture. After rapid warming (control), survival was high with EFS30 (79–93%) and EFS40 (96–99%). After slow warming, survival decreased with both EFS30 (48–62%) and EFS40 (44–64%). This must be from the formation of ICI. To examine the temperature at which ICI formed during slow warming, vitrified embryos were kept at various sub-zero temperatures during warming. Survival with EFS30 and EFS40 decreased on keeping samples for 3 min at −80 (25–75%), −60 (7–49%), −40 (0–41%), or −20 °C (26–60%). When samples were kept at −80 °C for 24 h, the survival decreased to 0–14%. These results suggest that ICI forms at a wide range of temperatures including −80 and −20 °C, more likely between −60 and −40 °C, and the ice forms not only quickly but also slowly.  相似文献   

14.
Conformational changes of apo A-1, the principal apoprotein of human plasma high density lipoprotein, have been studied by differential scanning calorimetry and ultraviolet difference spectroscopy as a function of temperature, pH, concentration of apoprotein, and urea concentration. Calorimetry shows that apo A-1 (5 to 40 mg/ml, pH 9.2) undergoes a two-state, reversible denaturation (enthalpy = 64 +/- 8.9 kcal/mole), between 43--71 degrees (midpoint temperature, Tm = 54 degrees), associated with a rise in heat capacity (deltaCvd) of 2.4 +/- 0.5 kcal/mole/degrees C. Apo A-1 (0.2 to 0.4 mg/ml, pH 9.2) develops a negative difference spectrum between 42--70 degrees, with Tm = 53 degrees. The enthalpy (deltaH = 59 +/- 5.7 kcal/mole at Tm) and heat capacity change (2.7 +/- 0.9 kcal/mole/degrees C) in the spectroscopic experiments were not significantly different from the calorimetric values. Below pH 9 and above pH 11, the calorimetric Tm and deltaH of denaturation are decreased. In the pH range of reversible denaturation (6.5 to 11.8), delatH and Tm are linearly related, showing that the heat capacity change (ddeltaH/dT) associated with denaturation is independent of Tm. In urea solutions, the calorimetric Tm and deltaH of denaturation are decreased. At 25 degrees, apo A-1 develops a negative difference spectrum between 1.4 and 3 M urea. Fifty per cent of the spectral change occurs in 2.4 M urea, which corresponds to the urea concentration obtained by extrapolation of the calorimetric Tm to 25 degrees. In urea solution of less than 0.75 M there is hyperchromicity at 285 nm (delta epsilon = 264 in 0.75 M urea), indicating strong interaction of aromatic amino acid residues in the native molecule with the solvent. Spectrophotometric titration of apo A-1 shows that 6.6 of the 7 tyrosine groups of apo A-1 titrate at pH less than 11.9, with similar titration curves obtained in aqueous solutions and in 6 M urea. The free energy of stabilization (deltaG) of the native conformation of apo A-1 was estimated, (a) at 37 degrees, using the calorimetric deltaA and deltaCvd, and (b) at 25 degrees, by extrapolation of spectroscopic data to zero urea concentration. The values (deltaG (37 degrees) = 2.4 and deltaG (25 degrees) = 2.7 kcal/mole) are small compared to typical globular proteins, indicating that native apo A-1 has a loosely folded tertiary structure. The low values of deltaG reflect the high degree of exposure of hydrophobic areas in the native protein molecule. The loosely folded conformation of apo A-1 allows extensive binding of lipid, since this can involve both surface hydrophobic sites and hydrophobic areas exposed by a cooperative, low energy unfolding process.  相似文献   

15.
Human immunodeficiency virus type 1 (HIV-1) genetic diversity poses a challenge to reliable viral load monitoring. Discrepancies between different testing platforms have been observed, especially for non-clade-B virus. Therefore we compare, in antiretroviral therapy (ART)-naïve South African subjects predominantly infected with HIV-1 clade-C, three commercially available assays: the COBAS AmpliPrep/COBAS TaqMan HIV-1 Test version 2.0 by Roche (CAP/CTM v2.0), the BioMérieux NucliSens Version 2.0 Easy Q/Easy Mag (NucliSens v2.0) and the Roche COBAS Amplicor HIV-1 Monitor Test Version 1.5 (Amplicor v1.5). Strong linear correlation was observed and Bland-Altman analyses showed overall good agreement between the assays with mean viral load differences of 0.078 log cp/ml (NucliSens v2.0 – Amplicor v1.5), 0.260 log cp/ml (CAP/CTM v2.0 – Amplicor v1.5) and 0.164 log cp/ml (CAP/CTM v2.0 – NucliSens v2.0), indicating lower mean viral load results for the Amplicor v1.5 and higher mean readings for the CAP/CTM v2.0. Consistent with observations following previous comparisons of CAP/CTM v2.0 versus Amplicor v1.5, the CAP/CTM v2.0 assay detected low-level viremia (median 65 cp/ml) in more than one-third of those in whom viremia had been undetectable (<20 cp/ml) in assays using the NucliSens platform. These levels of viremia are of uncertain clinical significance but may be of importance in early detection of ART resistance in those on treatment. Overall the three assays showed good comparability of results but with consistent, albeit relatively small, discrepancies for HIV-1 clade-C samples, especially in the low-viremic range that should be taken into account when interpreting viral load data.  相似文献   

16.
Heat effects of polyG-polyC melting in neutral aqueous solutions have been measured using differential scanning microcalorimeter with an extended temperature range. The limiting value of melting enthalpy is 53 +/- 4 kJ per mole of base pairs and melting temperature dependence on the sodium concentration can be expressed by the empiric relation Tm = 13.2 log(Na+) + 420 K.  相似文献   

17.
Spin labeled nucleic acids.   总被引:2,自引:2,他引:0       下载免费PDF全文
Homopolyribonucleotides and E. coli DNA wer spin labeled with an iodoacetamide-nitroxide compound. The extent of labeling is highly dependent upon the nature of the base and the secondary structure of the nucleic acid. This spin label-polymer linkage is unstable at high temperatures and in phosphate buffers. In order to determine the effect of changes in the environment of nucleic acids on the esr signals of their attached spin labels, the polynucleotides were subjected to temperature and viscosity perturbations. An increase in temperature (T) affects a linear decrease in the anisotropy factor of the esr signal. The log tau (tau = correlation time) versus (1/T) profile is linear with a positive slope when the spin label is attached to single stranded polynucleotides but exhibits discontinuities at certain critical temperatures when attached to the duplexes poly (As-U) and poly (I-Cs). These critical temperatures are lower than the optical Tm. Logarithmic increase in viscosity was found to produce a linear increase in tau in aqueous sucrose solutions.  相似文献   

18.
Climate warming and associated sea ice reductions in Antarctica have modified habitat conditions for some species. These include the congeneric Adélie, chinstrap and gentoo penguins, which now demonstrate remarkable population responses to regional warming. However, inconsistencies in the direction of population changes between species at different study sites complicate the understanding of causal processes. Here, we show that at the South Orkney Islands where the three species breed sympatrically, the less ice‐adapted gentoo penguins increased significantly in numbers over the last 26 years, whereas chinstrap and Adélie penguins both declined. These trends occurred in parallel with regional long‐term warming and significant reduction in sea ice extent. Periodical warm events, with teleconnections to the tropical Pacific, caused cycles in sea ice leading to reduced prey biomass, and simultaneous interannual population decreases in the three penguin species. With the loss of sea ice, Adélie penguins were less buffered against the environment, their numbers fluctuated greatly and their population response was strong and linear. Chinstrap penguins, considered to be better adapted to ice‐free conditions, were affected by discrete events of locally increased ice cover, but showed less variable, nonlinear responses to sea ice loss. Gentoo penguins were temporarily affected by negative anomalies in regional sea ice, but persistent sea ice reductions were likely to increase their available niche, which is likely to be substantially segregated from that of their more abundant congeners. Thus, the regional consequences of global climate perturbations on the sea ice phenology affect the marine ecosystem, with repercussions for penguin food supply and competition for resources. Ultimately, variability in penguin populations with warming reflects the local balance between penguin adaptation to ice conditions and trophic‐mediated changes cascading from global climate forcing.  相似文献   

19.
Lyophilization is the most popular method for achieving improved stability of labile biopharmaceuticals, but a significant fraction of product activity can be lost during processing due to stresses that occur in both the freezing and the drying stages. The effect of the freezing rate on the recovery of herpes simplex virus 2 (HSV-2) infectivity in the presence of varying concentrations of cryoprotectant excipients is reported here. The freezing conditions investigated were shelf cooling (223 K), quenching into slush nitrogen (SN2), and plunging into melting propane cooled in liquid nitrogen (LN2). The corresponding freezing rates were measured, and the ice crystal sizes formed within the samples were determined using scanning electron microscopy (SEM). The viral activity assay demonstrated the highest viral titer recovery for nitrogen cooling in the presence of low (0.25% w/v sucrose) excipient concentration. The loss of viral titer in the sample cooled by melting propane was consistently the highest among those results from the alternative cooling methods. However, this loss could be minimized by lyophilization at lower temperature and higher vacuum conditions. We suggest that this is due to a higher ratio of ice recrystallization for the sample cooled by melting propane during warming to the temperature at which freeze-drying was carried out, as smaller ice crystals readily enlarge during warming. Under the same freezing condition, a higher viral titer recovery was obtained with a formulation containing a higher concentration of sugar excipients. The reason was thought to be twofold. First, sugars stabilize membranes and proteins by hydrogen bonding to the polar residues of the biomolecules, working as a water substitute. Second, the concentrated sugar solution lowers the nucleation temperature of the water inside the virus membrane and prevents large ice crystal formation within both the virus and the external medium.  相似文献   

20.
P Boutron 《Cryobiology》1986,23(1):88-102
The glass-forming tendency and stability of the wholly amorphous state of various cryoprotective solutions has been studied in recent years (5-10, 20). A lot of experimental data including heats of ice crystallization at various cooling rates and devitrification temperatures have been given. In this article these data have been compared with analytical expressions using a semiempirical model. The theoretical variation of the total quantity of ice crystallized with the cooling rate fits very well with the experimental data, adjusting only one parameter. Using the same model, theoretical differential scanning calorimeter (DSC) crystallization peaks have been obtained for cooling or rewarming. The general shape, height, and width of the theoretical peaks are very similar to those of the experimental peaks. The differences are comparable to the random variations of the experimental peaks from one experiment to another. The analytical expressions obtained here could be used to study the relationship between the kinetics of ice crystallization and cell damage when ice crystallizes incompletely inside or outside the cells. These expressions have been applied to ice crystallization for applications in cryobiology. But they could also probably be used in other fields of research such as crystallization from silicates or other mineral or organic glasses.  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