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.
A method using magnesium oxide crystals to label the surface of physical gels, such as gelatin gel before quick-freezing is described and discussed. The quick-freeze, deep-etch, rotary replication technique is most adapted to 3-D visualization of physical gel structure. However, it is known that the depth which ultrarapid freezing may reach is limited by the growth of ice crystals as the distance from the surface of the specimen (rapidly cooled by smashing against a cooled metal plate) increases. Consequently, intact preservation of structures occurs only in superficial zones of the specimen. The MgO surface labelling technique provides a simple means for surface recognition. It enables the estimation of a given replicated area depth, taking into account the angle of specimen scraping before etching and replicating. By comparison of views of the same replica at different depths, freezing artifacts may be recognized even when they cause only slight deformations in the structure. This is particularly necessary for interpretation of gel network geometry: interpretation can be made with certainty only if a reliable surface reference marker exists. For gelatin gels, the depth of best freezing can be estimated to be around 5 μm from the frozen sample surface.  相似文献   

2.
Ice fraction was measured for solutions containing glucose, sucrose, gelatin, and egg albumin at various concentrations at temperatures from 0 to -20°C. For glucose and sucrose solutions, the ice fraction was accurately measured from phase diagram, which could be interpreted by solution thermodynamics with two parameters. The ice fractions of these sample solutions increased with decreases in both temperature and concentration. Because of the limited applicability of the phase diagram method only to systems with low molecular weight materials, the DSC method was also used for ice fraction measurement. The DSC method, corrected for temperature-dependent latent heat of ice and corrected with Pham’s equation, provided a good approximation for ice fractions with general applicability. The DSC method was used to measure the ice fractions of gelatin and egg albumin gels as a function of solute concentration. The freezing point and bound water of gelatin and egg albumin gels were described as a function of concentration. Effects of the differences in molecular structure on ice fraction were analyzed for various carbohydrate solutions at the same concentration. The ice fraction proved to be strongly dependent on the colligative properties of the solution with nonideal behavior.  相似文献   

3.
A differential scanning calorimeter has been developed for the automatic detection and measurement of dropwise freezing within a sample of 100-200 water drops. A typical drop size of 1 microliter is employed. The sample is distributed on flat, square (4-cm) thermoelectric sensors and the temperature is scanned downward by conductive cooling to a liquid nitrogen bath. The rate of cooling, typically 1 degree C/min, is set by the choice of a heat conduction rod between the calorimeter and the liquid nitrogen bath. The voltages from the thermopiles along with a system temperature-measuring thermocouple are continuously monitored by digital voltmeters and recorded every half-second in a computer memory. A freezing event in a drop is detected by a characteristic voltage signal whose integral with time is proportional to the size of the drop and its heat of fusion. The half-life of a freezing event signal is 10 s for a 1-microliter drop. The integrated signal produced from multiple freezing events is shown to provide a direct measure of the number of drops frozen at a given temperature. A distribution curve and its smoothed derivative can be constructed directly from these measurements. The instrument, which is termed an "ice nucleometer," is illustrated in determining the ice nucleation distribution in a population of Escherichia coli harboring cloned ice nucleation genes.  相似文献   

4.
Plane front freezing presents the possibility of encapsulating individual cells in the ice phase. The cells may also be pushed ahead of the plane front ice interface, as is always the case for conventional dendritic freezing, where the cells are pushed ahead of the thickening dendrite arms. Cells which are encapsulated during freezing are exposed to hypotonic liquid (pure water) initially upon thawing, while cells which are pushed into the last liquid to freeze are exposed to hypertonic liquid upon thawing. Some exposure to hypertonic intercellular liquid prior to freezing may be required to build up the salt and CPA content in the intracellular liquid and thereby avoid intracellular ice formation at the given cooling rate. Encapsulation of cells by a plane front ice interface should result in three regions of cell survival in the sample: an initial region of cell death due to intracellular ice formation, a final region of cell death due to overexposure to hypertonic intercellular liquid, and an intermediate region of cell survival, where neither damage mechanism has operated to a lethal level. An advantage of plane front freezing over dendritic freezing is that the regions of cell survival and death should be geometrically separate in the sample, rather than mixed at the dendritic microstructural level, as is the case for dendritic freezing. Samples containing populations with very high or very low survival rates for spermatozoa could be obtained by simply cutting up the frozen sample.  相似文献   

5.
Porous gelatin hydrogels: 1. Cryogenic formation and structure analysis   总被引:1,自引:0,他引:1  
In the present work, porous gelatin scaffolds were prepared by cryogenic treatment of a chemically cross-linked gelatin hydrogel, followed by removal of the ice crystals formed through lyophilization. This technique often leads to porous gels with a less porous skin. A simple method has been developed to solve this problem. The present study demonstrates that the hydrogel pore size decreased with an increasing gelatin concentration and with an increasing cooling rate of the gelatin hydrogel. Variation of the cryogenic parameters applied also enabled us to develop scaffolds with different pore morphologies (spherical versus transversal channel-like pores). In our opinion, this is the first paper in which temperature gradients during controlled cryogenic treatment were applied to induce a pore size gradient in gelatin hydrogels. With a newly designed cryo-unit, temperature gradients of 10 and 30 degrees C were implemented during the freezing step, resulting in scaffolds with average pore diameters of, respectively, +/-116 and +/-330 microm. In both cases, the porosity and pore size decreased gradually through the scaffolds. Pore size and structure analysis of the matrices was accomplished through a combination of microcomputed tomography using different software packages (microCTanalySIS and Octopus), scanning electron microscopy analysis, and helium pycnometry.  相似文献   

6.
To estimate theoretically how suited different freezing techniques are for freezing of freeze-etch specimens, it is necessary to know the relationship between specimen cooling rate and the resulting average ice crystal size. Using a somewhat simplified theoretical analysis, we have derived the approximate ice crystal size distribution of nonvitrified frozen aqueous specimens frozen at different cooling rates. The derived size distribution was used to calculate the relationship between relative change in average ice crystal size, (delta l/l), and relative change in specimen cooling rate delta (dT/dt)/(dT/dt). We found this relationship to be (delta l/l) = -k X delta (dT/dt)/(dT/dt) where k = 1.0 when specimen solidification takes place at about -6 degrees C, and k congruent to 1.3 when it takes place at about -40 degrees C.  相似文献   

7.
We have carried out a theoretical analysis of specimen cooling rate under ideal conditions during impact freezing and liquid-jet freezing. The analysis shows that use of liquid helium instead of liquid nitrogen as cooling medium during impact freezing results in an increase in a specimen cooling rate of no more than 30-40%. We have further shown that when both impact freezing and liquid-jet freezing are conducted at liquid nitrogen temperature, the two methods give approximately the same specimen cooling rate under ideal conditions except for a thin outer layer of the specimen. In this region impact freezing yields the highest cooling rate.  相似文献   

8.
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   

9.
The rate of ice formation was measured for Hedera helix L. cv. Thorndale (English ivy) bark exposed to -10 C. The cooling rate of bark exposed to -10 C was 31 C per minute. The water efflux rate required for ice formation to occur extracellularly was calculated from the rate of ice formation and the average cell diameter. The water potential difference driving the efflux of water to sites of extracellular ice was calculated from the sample temperature, osmotic water potential, and fraction of water frozen at a given freezing temperature. From the water efflux rate and water potential difference, the resistance of the barrier controlling movement of intracellular water to sites of extracellular ice was calculated. Comparison of the resistance of this barrier to water movement with the resistance of the cell membrane revealed that the membrane represented only 0.5% of the barrier resistance. Thus, membrane resistance can have little influence on the rate of water efflux and ice formation when bark is cooled at a rate of 31 C per minute. If ice formation occurred at the same rate in ivy bark as it occurred in a 10 mm MnCl(2) solution, the membrane resistance would still have represented only 1% of the resistance of the barrier to ice formation. Therefore, at a cooling rate of 31 C/minute, heat removal plays a large part in determining the rate of ice formation. At slower cooling rates experienced under natural freezing conditions the ability to remove heat would play an even larger role. It is concluded that under natural freezing conditions membrane resistance does not limit water efflux.  相似文献   

10.
Previous studies of the mechanism of freezing of erythrocytes in the absence of intracellular ice have been extended to define the catalytic sites responsible for promoting nucleation. The following aspects have been investigated: (1) the freeze propagation between undercooled erythrocytes, (2) the nucleation of ice in undercooled erythrocyte ghosts, and (3) the freezing behavior of undercooled hemoglobin solutions. The main findings are: (1) no cross-nucleation occurs between individual cells packed within the same emulsified water droplet; (2) the differential scanning calorimetric power-time curves of intact cells and ghosts are identical, indicating that hemoglobin does not affect ice nucleation; (3) the nucleation temperature of ice in an aqueous solution of hemoglobin (isolated from the cells) is substantially lower than that for the same solution when contained in the intact cell; (4) the threefold freeze concentration which accompanies the freezing of a 25% hemoglobin solution does not cause denaturation of the protein.  相似文献   

11.
Artificial gels, composed of collagen with or without hyaluronate (HA), a glycosaminoglycan (GAG), and chondroitin sulfate (CS), were prepared and quick-frozen for the purpose of studying the influence of composition and concentration on ice patterns. Dilute gels were spread on coverslips, plunged into a slush of 30% isopentane/70% propane (-185 degrees C), freeze-substituted, and examined by phase-contrast microscopy. Ice patterns were revealed as "ice cavities" in the gel after freeze-substitution. Ice morphology in the gels was gel-type-specific, suggesting that composition in dilute gels can influence ice pattern formation. Crystallization patterns reflecting high, intermediate, and low rates of freezing were observed in all gel types. Intermediate freezing in differentiating gel-type-specific ice patterns. Gels which included hyaluronate (HA) and chondroitin sulfate (CS) altered the ice crystal pattern commonly observed in collagen gels. Ice structure in collagen gels consisted predominantly of long, parallel crystals in the herringbone pattern. Ice crystals separated gel into thin, unbranched fibers with a primary spacing of approximately 2 microns. Ice morphology in HA gels formed a mosaic consisting of packets of ice crystals. Contiguous packets were often oriented at right angles to each other. Periodic crossbridges interconnect primary gel fibers of HA gels and interrupt the lengthwise growth of ice crystals. Smooth beads were visible on primary strands in HA gels frozen at intermediate velocities. The addition of CS to collagen gels resulted in formation of randomly oriented ice crystals in gels frozen at intermediate rates. CS has little influence on ice morphology at low freezing velocities. Primary strands in CS gels were decorated with rough-surfaced, osmiophilic aggregates.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
M S Brown  F W Reuter 《Cryobiology》1974,11(3):185-191
A new technique was developed for the observation and recording on videotape of thermal and microscopic changes that occur simultaneously during the freezing of cucumber tissue. The freezing process occurs in two steps. Nucleation and growth of ice crystals in the continuous extracellular liquid phase is followed by nucleation and growth of ice crystals in individual supercooled cells. The freezing of cells in rapid succession causes the average temperature to remain constant for a short time. This mechanism explains the second freezing plateau found in most plant tissue freezing curves.  相似文献   

13.
《Cryobiology》2006,52(3):262-280
Antifreeze proteins are characterised by their ability to prevent ice from growing upon cooling below the bulk melting point. This displacement of the freezing temperature of ice is limited and at a sufficiently low temperature a rapid ice growth takes place. The separation of the melting and freezing temperature is usually referred to as thermal hysteresis, and the temperature of ice growth is referred to as the hysteresis freezing point. The hysteresis is supposed to be the result of an adsorption of antifreeze proteins to the crystal surface. This causes the ice to grow as convex surface regions between adjacent adsorbed antifreeze proteins, thus lowering the temperature at which the crystal can visibly expand. The model requires that the antifreeze proteins are irreversibly adsorbed onto the ice surface within the hysteresis gap. This presupposition is apparently in conflict with several characteristic features of the phenomenon; the absence of superheating of ice in the presence of antifreeze proteins, the dependence of the hysteresis activity on the concentration of antifreeze proteins and the different capacities of different types of antifreeze proteins to cause thermal hysteresis at equimolar concentrations. In addition, there are structural obstacles that apparently would preclude irreversible adsorption of the antifreeze proteins to the ice surface; the bond strength necessary for irreversible adsorption and the absence of a clearly defined surface to which the antifreeze proteins may adsorb. This article deals with these apparent conflicts between the prevailing theory and the empirical observations. We first review the mechanism of thermal hysteresis with some modifications: we explain the hysteresis as a result of vapour pressure equilibrium between the ice surface and the ambient fluid fraction within the hysteresis gap due to a pressure build-up within the convex growth zones, and the ice growth as the result of an ice surface nucleation event at the hysteresis freezing point. We then go on to summarise the empirical data to show that the dependence of the hysteresis on the concentration of antifreeze proteins arises from an equilibrium exchange of antifreeze proteins between ice and solution at the melting point. This reversible association between antifreeze proteins and the ice is followed by an irreversible adsorption of the antifreeze proteins onto a newly formed crystal plane when the temperature is lowered below the melting point. The formation of the crystal plane is due to a solidification of the interfacial region, and the necessary bond strength is provided by the protein “freezing” to the surface. In essence: the antifreeze proteins are “melted off” the ice at the bulk melting point and “freeze” to the ice as the temperature is reduced to subfreezing temperatures. We explain the different hysteresis activities caused by different types of antifreeze proteins at equimolar concentrations as a consequence of their solubility features during the phase of reversible association between the proteins and the ice, i.e., at the melting point; a low water solubility results in a large fraction of the proteins being associated with the ice at the melting point. This leads to a greater density of irreversibly adsorbed antifreeze proteins at the ice surface when the temperature drops, and thus to a greater hysteresis activity. Reference is also made to observations on insect antifreeze proteins to emphasise the general validity of this approach.  相似文献   

14.
The mechanism by which fish antifreeze proteins cause thermal hysteresis   总被引:6,自引:0,他引:6  
Antifreeze proteins are characterised by their ability to prevent ice from growing upon cooling below the bulk melting point. This displacement of the freezing temperature of ice is limited and at a sufficiently low temperature a rapid ice growth takes place. The separation of the melting and freezing temperature is usually referred to as thermal hysteresis, and the temperature of ice growth is referred to as the hysteresis freezing point. The hysteresis is supposed to be the result of an adsorption of antifreeze proteins to the crystal surface. This causes the ice to grow as convex surface regions between adjacent adsorbed antifreeze proteins, thus lowering the temperature at which the crystal can visibly expand. The model requires that the antifreeze proteins are irreversibly adsorbed onto the ice surface within the hysteresis gap. This presupposition is apparently in conflict with several characteristic features of the phenomenon; the absence of superheating of ice in the presence of antifreeze proteins, the dependence of the hysteresis activity on the concentration of antifreeze proteins and the different capacities of different types of antifreeze proteins to cause thermal hysteresis at equimolar concentrations. In addition, there are structural obstacles that apparently would preclude irreversible adsorption of the antifreeze proteins to the ice surface; the bond strength necessary for irreversible adsorption and the absence of a clearly defined surface to which the antifreeze proteins may adsorb. This article deals with these apparent conflicts between the prevailing theory and the empirical observations. We first review the mechanism of thermal hysteresis with some modifications: we explain the hysteresis as a result of vapour pressure equilibrium between the ice surface and the ambient fluid fraction within the hysteresis gap due to a pressure build-up within the convex growth zones, and the ice growth as the result of an ice surface nucleation event at the hysteresis freezing point. We then go on to summarise the empirical data to show that the dependence of the hysteresis on the concentration of antifreeze proteins arises from an equilibrium exchange of antifreeze proteins between ice and solution at the melting point. This reversible association between antifreeze proteins and the ice is followed by an irreversible adsorption of the antifreeze proteins onto a newly formed crystal plane when the temperature is lowered below the melting point. The formation of the crystal plane is due to a solidification of the interfacial region, and the necessary bond strength is provided by the protein "freezing" to the surface. In essence: the antifreeze proteins are "melted off" the ice at the bulk melting point and "freeze" to the ice as the temperature is reduced to subfreezing temperatures. We explain the different hysteresis activities caused by different types of antifreeze proteins at equimolar concentrations as a consequence of their solubility features during the phase of reversible association between the proteins and the ice, i.e., at the melting point; a low water solubility results in a large fraction of the proteins being associated with the ice at the melting point. This leads to a greater density of irreversibly adsorbed antifreeze proteins at the ice surface when the temperature drops, and thus to a greater hysteresis activity. Reference is also made to observations on insect antifreeze proteins to emphasise the general validity of this approach.  相似文献   

15.
Fourier transform infrared spectroscopy (FTIR) and cryomicroscopy were used to define the process of cellular injury during freezing in LNCaP prostate tumor cells, at the molecular level. Cell pellets were monitored during cooling at 2 degrees C/min while the ice nucleation temperature was varied between -3 and -10 degrees C. We show that the cells tend to dehydrate precipitously after nucleation unless intracellular ice formation occurs. The predicted incidence of intracellular ice formation rapidly increases at ice nucleation temperatures below -4 degrees C and cell survival exhibits an optimum at a nucleation temperature of -6 degrees C. The ice nucleation temperature was found to have a great effect on the membrane phase behavior of the cells. The onset of the liquid crystalline to gel phase transition coincided with the ice nucleation temperature. In addition, nucleation at -3 degrees C resulted in a much more co-operative phase transition and a concomitantly lower residual conformational disorder of the membranes in the frozen state compared to samples that nucleated at -10 degrees C. These observations were explained by the effect of the nucleation temperature on the extent of cellular dehydration and intracellular ice formation. Amide-III band analysis revealed that proteins are relatively stable during freezing and that heat-induced protein denaturation coincides with an abrupt decrease in alpha-helical structures and a concomitant increase in beta-sheet structures starting at an onset temperature of approximately 48 degrees C.  相似文献   

16.
It was found that freezing of water in terms of homogeneous nucleation of ice never occurs even in ultra-clean micro-sized water droplets under normal conditions. More surprisingly, at sufficiently low supercoolings, foreign nano-particles exert no effect on the nucleation barrier of ice; it is as if they physically "vanished." This effect, called hereafter the "zero-sized" effect of foreign particles (or nucleators), leads to the entry of a so-called inverse homogeneous-like nucleation domain, in which nucleation is effectively suppressed. The freezing temperature of water corresponds to the transition temperature from the inverse homogeneous-like nucleation regime to foreign particle-mediated heterogeneous nucleation. The freezing temperature of water is mainly determined by (i) the surface roughness of nucleators at large supercoolings, (ii) the interaction and structural match between nucleating ice and the substrate, and (iii) the size of the effective surface of nucleators at low supercoolings. Our experiments showed that the temperature of -40 degrees C, commonly regarded as the temperature of homogeneous nucleation-mediated freezing, is actually the transition temperature from the inverse homogeneous-like nucleation regime to foreign particle-mediated heterogeneous nucleation in ultra-clean water. Taking advantage of inverse homogeneous-like nucleation, the interfacial tensions between water and ice in very pure water and antifreeze aqueous solutions were measured at a very high precision for the first time. The principles of freezing promotion and antifreeze and the selection for the biological ice nucleation and antifreeze proteins are obtained. The results provide completely new insights into freezing and antifreeze phenomena and bear generic implications for all crystallization systems.  相似文献   

17.
Using cryo‐SEM with EDX fundamental structural and mechanical properties of the moss Ceratodon purpureus (Hedw.) Brid. were studied in relation to tolerance of freezing temperatures. In contrast to more complex plants, no ice accumulated within the moss during the freezing event. External ice induced desiccation with the response being a function of cell type; water‐filled hydroid cells cavitated and were embolized at ?4 °C while parenchyma cells of the inner cortex exhibited cytorrhysis, decreasing to ~20% of their original volume at a nadir temperature of ?20 °C. Chlorophyll fluorescence showed that these winter acclimated mosses displayed no evidence of damage after thawing from ?20 °C while GCMS showed that sugar concentrations were not sufficient to confer this level of freezing tolerance. In addition, differential scanning calorimetry showed internal ice nucleation occurred in hydrated moss at ~?12 °C while desiccated moss showed no evidence of freezing with lowering of nadir temperature to ?20 °C. Therefore the rapid dehydration of the moss provides an elegantly simple solution to the problem of freezing; remove that which freezes.  相似文献   

18.
《Biophysical journal》2020,118(3):708-719
Blotting has been the standard technique for preparing aqueous samples for single-particle electron cryo-microscopy for over three decades. This technique removes the excess solution from a transmission electron microscope grid by pressing absorbent filter paper against the specimen before vitrification. However, this standard technique produces vitreous ice with inconsistent thickness from specimen to specimen and from region to region within the same specimen, the reasons for which are not understood. Here, high-speed interference contrast microscopy is used to demonstrate that the irregular pattern of fibers in the filter paper imposes tortuous, highly variable boundaries during the removal of excess liquid from a flat, hydrophilic surface. As a result, aqueous films of nonuniform thickness are formed while the filter paper is pressed against the substrate. This pattern of nonuniform liquid thickness changes again after the filter paper is pulled away, but the thickness still does not become completely uniform. We suggest that similar topographical features of the liquid film are produced during the standard technique used to blot EM grids and that these manifest in nonuniform ice after vitrification. These observations suggest that alternative thinning techniques, which do not rely on direct contact between the filter paper and the grid, may result in more repeatable and uniform sample thicknesses.  相似文献   

19.
Calorimetric analysis indicates that 82% of the body water of Hemideina maori is converted into ice at 10 degrees C. This is a high proportion and led us to investigate whether intracellular freezing occurs in H. maori tissue. Malpighian tubules and fat bodies were frozen in haemolymph on a microscope cold stage. No fat body cells, and 2% of Malpighian tubule cells froze during cooling to -8 degrees C. Unfrozen cells appeared shrunken after ice formed in the extracellular medium. There was no difference between the survival of control tissues and those frozen to -8 degrees C. At temperatures below -15 degrees C (lethal temperatures for weta), there was a decline in survival, which was strongly correlated with temperature, but no change in the appearance of tissue. It is concluded that intracellular freezing is avoided by Hemideina maori through osmotic dehydration and freeze concentration effects, but the reasons for low temperature mortality remain unclear. The freezing process in H. maori appears to rely on extracellular ice nucleation, possibly with the aid of an ice nucleating protein, to osmotically dehydrate the cells and avoid intracellular freezing. The lower lethal temperature of H. maori (-10 degrees C) is high compared to organisms that survive intracellular freezing. This suggests that the category of 'freezing tolerance' is an oversimplification, and that it may encompass at least two strategies: intracellular freezing tolerance and avoidance.  相似文献   

20.
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   

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