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 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   

2.
Intragastric administration of certain salt components of mineral water naftusia in minimal concentrations (1-5 mmol/l) changes intensity and direction of hypergastrinemic reaction in the rats. The effect is determined by the anionic composition of salts rather than by the cationic one and depends on the initial concentration of gastrin in blood.  相似文献   

3.
Mono- and biphasic kinetic effects of bile salts on the pancreatic IB phospholipase A2 (PLA2) catalyzed interfacial hydrolysis are characterized. This novel phenomenon is modeled as allosteric action of bile salts with PLA2 at the interface. The results and controls also show that these kinetic effects are not due to surface dilution or solubilization or disruption of the bilayer interface where in the mixed-micelles substrate replenishment becomes the rate-limiting step. The PLA2-catalyzed rate of hydrolysis of zwitterionic dimyristoylphosphatidylcholine (DMPC) vesicles depends on the concentration and structure of the bile salt. The sigmoidal rate increase with cholate saturates at 0.06 mole fraction and changes little at the higher mole fractions. Also, with the rate-lowering bile salts (B), such as taurochenodeoxycholate (TCDOC), the initial sigmoidal rate increase at lower mole fraction is followed by nearly complete reversal to the rate at the pre-activation level at higher mole fractions. The rate-lowering effect of TCDOC is not observed with the (62-66)-loop deleted DeltaPLA2, or with the Naja venom PLA2 that is evolutionarily devoid of the loop. The rate increase is modeled with the assumption that the binding of PLA2 to DMPC interface is cooperatively promoted by bile salt followed by allosteric k(cat)(*)-activation of the bound enzyme by the anionic interface. The rate-lowering effect of bile salts is attributed to the formation of a specific catalytically inert E(*)B complex in the interface, which is noticeably different than the 1:1 EB complex in the aqueous phase. The cholate-activated rate of hydrolysis is lowered by hypolidemic ezetimibe and guggul extract which are not interfacial competitive inhibitors of PLA2. We propose that the biphasic modulation of the pancreatic PLA2 activity by bile salts regulates gastrointestinal fat metabolism and cholesterol homeostasis.  相似文献   

4.
The biliary elimination of glycodihydrofusidate (GDHF), a structural analogue of bile salts, was studied in bile fistula rats. GDHF was excreted in bile with a maximal excretory rate (Tm = 0.80 mumol min-1 kg-1) which is much lower than bile salts Tm. The effects of dehydrocholate and taurocholate on GDHF biliary secretion suggest a stimulatory effect of bile salts on canalicular excretion of the drug. (a) When a bolus intravenous injection of 3 mumol of GDHF was followed after 2 min by a continuous dehydrocholate perfusion (10 mumol min-1 kg-1), biliary excretion of GDHF was increased in comparison with control rats. (b) Upon attaining the biliary Tm by continuous perfusion of GDHF at a rate of 1.35 mumol min-1 kg-1, infusion with either taurocholate or dehydrocholate increased its Tm to a similar degree. These results are similar to those previously obtained with the effects of bile salt infusions on the Tm of bromosulfophthalein. They suggest therefore that hepatic transport of GDHF and bile salts occurs by routes which are distinct for canalicular transport in spite of the striking structural similarities between GDHF and bile salts.  相似文献   

5.
Conditions for the in vitro activation of the latent form of a protease-like permeability factor in the pseudoglobulin fraction from guinea pig skin were examined. (1) The factor was activated by dialysis against 67 mM phosphate buffer at pH 5.8--6.4, not at pH 7.0--8.0. (2) High salt concentration (200 mM or greater phosphate buffer or 67 mM phosphate buffer containing 200 mM or greater KCl or NaCl) prevented the activation at pH 6.2. (3) High osmotic pressure (sucrose at 1 M) did not affect activation at pH 6.2. (4) Reconversion of the activated permeability factor into an inactive form was not observed under high salt conditions, under which the latent permeability factor was stable in its own form. (5) The molecular size of the latent permeability factor was estimated as approx. 80 000 by Sephadex G-100 gel filtration at high salt concentration.  相似文献   

6.
Covalent binding stoichiometries for both the enzyme:5-fluoro-2'-deoxyuridine 5'-monophosphate (FdUMP) binary complex and the enzyme:FdUMP:5,10-methylenetetrahydrofolate (inhibitory ternary) complex at equilibrium were measured by the trichloroacetic acid precipitation assay and shown to be a function of temperature, time, pH, salt concentration, buffer composition and thiol concentration. Incubation at 37 degrees C yielded the maximum covalent binding ratio (mol FdUMP/mol enzyme) for the latter binary (0.7) and ternary (1.7) complexes. In most buffers studied, the maximum covalent binding ratio (1.5-1.7) for the inhibitory ternary complex occurred over a broad pH range (4.5-8.0), while the optimum covalent binding ratio for binary complex was observed at a much narrower region centered between pH 5.5-6.5. In the presence of increasing concentrations of phosphate buffer, the maximum binding ratio for the covalent binary complex decreased from 0.63 in the absence of phosphate to 0.1 in the presence of 225 mM phosphate, while that for the inhibitory ternary complex was unchanged. When a ternary complex was formed with enzyme, FdUMP and (+/-)-tetrahydrofolate in the absence of phosphate, the FdUMP:enzyme covalent binding ratio was 1.8, while in the presence of 75 mM phosphate, the binding ratio was only 1.0. When exogenous thiol was removed by centrifugal column chromatography, the maximum binding stoichiometry of the resulting inhibitory ternary complex was 1.7 and was independent of added thiol over a 2 h incubation period at 37 degrees C. When extensive dialysis at 5 degrees C was used to remove the thiol, the maximum binding stoichiometry of the resulting inhibitory ternary complex was found to be dependent on both the concentration of added thiol and the time of incubation at 37 degrees C and did not exceed a value of 1.0.  相似文献   

7.
Mono- and biphasic kinetic effects of bile salts on the pancreatic IB phospholipase A2 (PLA2) catalyzed interfacial hydrolysis are characterized. This novel phenomenon is modeled as allosteric action of bile salts with PLA2 at the interface. The results and controls also show that these kinetic effects are not due to surface dilution or solubilization or disruption of the bilayer interface where in the mixed-micelles substrate replenishment becomes the rate-limiting step. The PLA2-catalyzed rate of hydrolysis of zwitterionic dimyristoylphosphatidylcholine (DMPC) vesicles depends on the concentration and structure of the bile salt. The sigmoidal rate increase with cholate saturates at 0.06 mole fraction and changes little at the higher mole fractions. Also, with the rate-lowering bile salts (B), such as taurochenodeoxycholate (TCDOC), the initial sigmoidal rate increase at lower mole fraction is followed by nearly complete reversal to the rate at the pre-activation level at higher mole fractions. The rate-lowering effect of TCDOC is not observed with the (62-66)-loop deleted ΔPLA2, or with the Naja venom PLA2 that is evolutionarily devoid of the loop. The rate increase is modeled with the assumption that the binding of PLA2 to DMPC interface is cooperatively promoted by bile salt followed by allosteric kcat?-activation of the bound enzyme by the anionic interface. The rate-lowering effect of bile salts is attributed to the formation of a specific catalytically inert E?B complex in the interface, which is noticeably different than the 1:1 EB complex in the aqueous phase. The cholate-activated rate of hydrolysis is lowered by hypolidemic ezetimibe and guggul extract which are not interfacial competitive inhibitors of PLA2. We propose that the biphasic modulation of the pancreatic PLA2 activity by bile salts regulates gastrointestinal fat metabolism and cholesterol homeostasis.  相似文献   

8.
The maximal equilibrium solubility of cholesterol in mixtures of phosphatidylcholine (PC)1 and bile salts depends on the cholesterol/PC ratio (Rc) and on the effective ratio (Re) between nonmonomeric bile salts and the sum (CT) of PC and cholesterol concentrations (Carey and Small, 1978; Lichtenberg et al., 1984). By contrast, the concentration of bile salts required for solubilization of liposomes made of PC and cholesterol does not depend on Rc (Lichtenberg et al., 1984 and 1988). Thus, for Rc greater than 0.4, solubilization of the PC-cholesterol liposomes yields PC-cholesterol-bile salts mixed micellar systems which are supersaturated with cholesterol. In these metastable systems, the mixed micelles spontaneously undergo partial revesiculation followed by crystallization of cholesterol. The rate of the latter processes depends upon Rc, Re, and CT. For any given Rc and Re, the rate of revesiculation increases dramatically with increasing the lipid concentration CT, reflecting the involvement of many mixed micelles in the formation of each vesicle. The rate also increases, for any given CT and Re, upon increasing the cholesterol to PC ratio, Rc, probably due to the increasing degree of supersaturation. Increasing the cholate to lipid effective ratio, Re, by elevation of cholate concentration at constant Rc and CT has a complex effect on the rate of the revesiculation process. As expected, cholate concentration higher than that required for complete solubilization at equilibrium yields stable mixed micellar systems which do not undergo revesiculation, but for lower cholate concentrations decreasing the degree of supersaturation (by increasing [cholate]) results in faster revesiculation. We interpret these results in terms of the structure of the mixed micelles; micelles with two or more PC molecules per one molecule of cholesterol are relatively stable but increasing the bile salt concentration may cause dissociation of such 1:2 cholesterol:PC complexes, hence reducing the stability of the mixed micellar dispersions. The instability of PC-cholesterol-cholate mixed systems with intermediary range of cholate to lipids ratio may be significant to gallbladder stone formation as: (a) biliary bile contains PC-cholesterol vesicles which may be, at least partially, solubilized by bile salts during the process of bile concentration in the gallbladder, resulting in mixtures similar to our model systems; and (b) the bile composition of cholesterol gallstone patients is within an intermediary range of bile salts to lipids ratio.  相似文献   

9.
The influence of sodium metaperiodate concentration on kinetics and conversion degree of peroxidase carbohydrate moiety as well as the effect of the oxidation degree of the carbohydrate moiety on the composition, structure and properties of insulin-peroxidase conjugates were studied. The initial rate of peroxidase's oxidation is directly proportional to the periodate concentration; the oxidation rate constant of peroxidase carbohydrate moiety is 1.23 x 10(-3) M-1 min-1. At the molar ratio of metaperiodate to peroxidase 150:1 or higher, the maximal quantity of aldehyde groups (62 +/- 2) in the peroxidase molecule is formed and the oxidation of each carbohydrate chain leads to the formation of eight aldehyde groups. The molecular mass composition of the insulin-peroxidase conjugates was studied by HPLC. The conjugates proved to be multicomponent mixtures of oligomers (53, 83, 128, 174, 268, 440 kD and higher). The insulin-peroxidase molar ratio in the fractions of the conjugates with molecular masses higher than 83 kD is 8:1. It was shown that the affinity of insulin-peroxidase conjugates to antibodies depends on the oxidation degree of peroxidase used for production of conjugates.  相似文献   

10.
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   

11.
Phase transition temperatures of the water-rich part of the H2O-NaCl-HES ternary system have been studied by means of differential thermal analysis. The experimental data indicate that the protective action of the macromolecular agent HES is different from that of DMSO and glycerol which seems essentially to be due to colligative properties. As expected, there is no evidence of a significant freezing-point depression caused by HES concentrated up to 40 wt%. The occurrence of a ternary eutectic point may therefore be excluded. Instead, an isothermal eutectic trough is observed. The extrapolated course of its projection onto the basal composition triangle indicates that a certain portion of water is absorbed by HES and kept from freezing, i.e., appears to be thermally inert within the range of temperatures studied. The protective action of HES against solution effects, therefore, is attributed to its water-absorptive capacity and kinetics instead of a postponement of lethal salt enrichment to lower temperatures as caused by DMSO or glycerol. Consequently, it is possible to determine a minimum initial HES concentration that completely prevents lethal salt enrichment during cooling. For the case of red blood cells, the derived algebraic expression yields an initial HES fraction of 11 wt% if the respective values are inserted.  相似文献   

12.
The nature of the electron transfer reaction between reducedplastocyanin and P700 oxidized by flash illumination was studiedin P700-enriched Triton subchloroplast fraction 1 particles.An addition of monovalent salts to the suspension at neutralpH increased the reaction rate at low concentrations (>20mM). Salts of divalent cations showed a similar effect at muchlower concentrations (>2 mM), This effect was not dependenton the concentration and the valence of anions. The increaseof rate at low salt concentrations was observed at pH's above5, but below pH 5 the rate was decreased by adding salts. Atabout pH 5, the rate was not affected by salts. Apart from thesesalt effects, the optimum pH for the reaction rate was observedbetween 5.5 and 6.5. The reduction rate depended sigmoidally on the added plastocyaninconcentration at pH 6.8 and 4. A Michaelis-Menten type relationshipwas observed at about pH 5. The half-saturation concentrationof plastocyanin became lower as the salt concentration increasedat pH 6.8, while it became higher by adding salt at pH 4. The effects of salts on the rate of electron donation from othermetalloproteins and artificial electron donors to P700 werealso studied. It is concluded, from the analysis with the Gouy-Chapmantheory, that the net charges on the electron donors and themembrane surfaces mainly determine the response of the P700reduction rate to salt addition. The salt addition changes mainlythe local concentration or accessibility of electron donorsto P700. (Received January 12, 1981; Accepted March 6, 1981)  相似文献   

13.
Debaryomyces hansenii was grown in YPD medium without or with 1.0 M NaCl or KCl. Respiration was higher with salt, but decreased if it was present during incubation. However, carbonylcyanide-3-chlorophenylhydrazone (CCCP) markedly increased respiration when salt was present during incubation. Salt also stimulated proton pumping that was partially inhibited by CCCP; this uncoupling of proton pumping may contribute to the increased respiratory rate. The ADP increase produced by CCCP in cells grown in NaCl was similar to that observed in cells incubated with or without salts. The alternative oxidase is not involved. Cells grown with salts showed increased levels of succinate and fumarate, and a decrease in isocitrate and malate. Undetectable levels of citrate and low-glutamate dehydrogenase activity were present only in NaCl cells. Both isocitrate dehydrogenase decreased, and isocitrate lyase and malate synthase increased. Glyoxylate did not increase, indicating an active metabolism of this intermediary. Higher phosphate levels were also found in the cells grown in salt. An activation of the glyoxylate cycle results from the salt stress, as well as an increased respiratory capacity, when cells are grown with salt, and a 'coupling' effect on respiration when incubated in the presence of salt.  相似文献   

14.
Some characteristics of protein precipitation by salts   总被引:1,自引:0,他引:1  
The solubilities of lysozyme, alpha-chymotrypsin and bovine serum albumin (BSA) were studied in aqueous electrolyte solution as a function of ionic strength, pH, the chemical nature of salt, and initial protein concentration. Compositions were measured for both the supernatant phase and the precipitate phase at 25 degrees C. Salts studied were sodium chloride, sodium sulfate, and sodium phosphate. For lysozyme, protein concentrations in supernatant and precipitate phases are independent of the initial protein concentration; solubility can be represented by the Cohn salting-out equation. Lysozyme has a minimum solubility around pH 10, close to its isoelectric point (pH 10.5). The effectiveness of the three salts studied for precipitation were in the sequence sulfate > phosphate > chloride, consistent with the Hofmeister series. However, for alpha-chymotrypsin and BSA, initial protein concentration affects the apparent equillibrium solubility. For these proteins, experimental results show that the compositions of the precipitate phase are also affected by the initial protein concentration. We define a distribution coefficient kappa(e) to represent the equilibrium ratio of the protein concentration in the supernatant phase to that in the precipitate phase. When the salt concentration is constant, the results show that, for lysozyme, the protein concentrations in both phases are independent of the initial protein concentrations, and thus kappa(e) is a constant. For alpha-chymotrypsin and BSA, their concentrations in both phases are nearly proportional to the initial protein concentrations, and therefore, for each protein, at constant salt concentration, the distribution coefficient kappa(e) is independent of the initial protein concentration. However, for both lysozyme and alpha-chymotrypsin, the distribution coefficient falls with increasing salt concentration. These results indicate that care must be used in the definition of solubility. Solubility is appropriate when the precipitate phase is pure, but when it is not, the distribution coefficient better describes the phase behavior. (c) 1992 John Wiley & Sons, Inc.  相似文献   

15.
The rate of calcium transport by sarcoplasmic reticulum vesicles from dog heart assayed at 25 degrees C, pH 7.0, in the presence of oxalate and a low free Ca2+ concentration (approx. 0.5 microM) was increased from 0.091 to 0.162 mumol . mg-1 . min-1 with 100 nM calmodulin, when the calcium-, calmodulin-dependent phosphorylation was carried out prior to the determination of calcium uptake in the presence of a higher concentration of free Ca2+ (preincubation with magnesium, ATP and 100 microM CaCl2; approx. 75 microM free Ca2+). Half-maximal activation of calcium uptake occurs under these conditions at 10-20 nM calmodulin. The rate of calcium-activated ATP hydrolysis by the Ca2+-, Mg2+-dependent transport ATPase of sarcoplasmic reticulum was increased by 100 nM calmodulin in parallel with the increase in calcium transport; calcium-independent ATP splitting was unaffected. The calcium-, calmodulin-dependent phosphorylation of sarcoplasmic reticulum, preincubated with approx. 75 microM Ca2+ and assayed at approx. 10 microM Ca2+ approaches maximally 3 nmol/mg protein, with a half-maximal activation at about 8 nM calmodulin; it is abolished by 0.5 mM trifluperazine. More than 90% of the incorporated [32P]phosphate is confined to a 9-11 kDa protein, which is also phosphorylated by the catalytic subunit of the cAMP-dependent protein kinase and most probably represents a subunit of phospholamban. The stimulatory effect of 100 nM calmodulin on the rate of calcium uptake assayed at 0.5 microM Ca2+ was smaller following preincubation of sarcoplasmic reticulum vesicles with calmodulin in the presence of approx. 75 microM Ca2+, but in the absence of ATP, and was associated with a significant degree of calmodulin-dependent phosphorylation. However, the stimulatory effect on calcium uptake and that on calmodulin-dependent phosphorylation were both absent after preincubation with calmodulin, without calcium and ATP, suggestive of a causal relationship between these processes.  相似文献   

16.
1. The nuclear-myofibrilar (800g pellet) fraction of the uterus from immature (22-23 days old) rats not exposed to oestrogen exhibits saturable binding of oestradiol. The nuclear binding capacity represents approximately 10% of that of the cytosol fraction (approx. 3.5 fmol/mug of DNA). The predominant part (0.3.5 fmol/mug of DNA) of the nuclear binind sites are present in the residual pellet after extraction with 0.5 M-KC1. 2. By using an exchange technique in vitro, determinations of the nuclear binding sites have been carried out after administration of 1 mug of oestradiol in vivo. Within 0.5h after the hormone injection, the concentration of nuclear bindng sites increased to approx. 0.4 fmol/mug of DNA in the 0.5 M-KC1-extractable fraction, and to approx. 1.2 fmol/mug of DNA in the residual fraction. Meanwhile the cytosol oestrogen-receptor concentration decreased to approx. 10% of its initial value. In the following period from 0.5 h after the oestradiol injection onwards, the concentration of nuclear oestrogen receptors decreased with halflife values of approx. 140 and 200 min for the KC1(0.5 M)-extractable and residual form respectively. At the same time, the cytosol receptor concentration increased to reach approx. 50% of the initial value by the 6h. This increase could not be blocked by cycloheximide. The initial concentration of cytosol receptor was restored approx. 11h after the injection and the increase during the 6-11h period was sensitive to cycloheximide inhibition, suggesting protein-synthesis-dependence of the process. 3. With the (more) physiological dose of oestradiol (0.1 mug), the decrease the cytosol receptor was only 50% by 4h and this was followed by a period (up to 12h after injection) during which the initial concentration was restored. During this period the increase of the receptor can be blocked by cycloheximide.  相似文献   

17.
1. Addition of succinate to valinomycin-treated mitochondria incubated in KCL causes a large electrolyte penetration. The process depends on a steady supply of energy and involves a continuous net extrusion of protons. Rates of respiration and of electrolyte penetration proceed in a parallel manner. 2. A passive penetration of K+ salt of permeant anions occurs in respiratory-inhibited mitochondria after addition of valinomycin. Addition of succinate at the end of the passive swelling starts an active extrusion of anions and cations with restoration of the initial volume. The shrinkage is accompanied by a slow reuptake of protons. The initiation of the active shrinkage correlates with the degree of stretching of the inner membrane. The extrusion of electrolytes is inhibited by nigericin, while it is only slightly sensitive to variations of the valinomycin concentration larger than two orders of magnitude. 3. Passive swelling and active shrinkage occurs also when K+ is replaced by a large variety of organic cations. The rate of organic cation penetration is enhanced by tetraphenylboron, while the rate of electrolyte extrusion is insensitive to variation of the tetraphenylboron concentration. 4. Active shrinkage, either with K+ or organic cation salts, is inhibited by weak acids. The phosphate inhibition is removed by SH inhibitors. The active shrinkage is also inhibited by mersalyl to an extent of about 60%. 5. Three models of active shrinkage are discussed: (a) mechanoprotein, (b) electrogenic proton pump, and (c) proton-driven cation anion pump.  相似文献   

18.
As suspensions of cells freeze, the electrolytes and other solutes in the external solution concentrate progressively, and the cells undergo osmotic dehydration if cooling is slow. The progressive concentration of solute comes about as increasing amounts of pure ice precipitate out of solution and cause the liquid-filled channels in which the cells are sequestered to dwindle in size. The consensus has been that slow freezing injury is related to the composition of the solution in these channels and not to the amount of residual liquid. The purpose of the research reported here was to test this assumption on human erythrocytes. Ordinarily, solute concentration and the amount of liquid in the unfrozen channels are inversely coupled. To vary them independently, one must vary the initial solute concentration. Two solutes were used here: NaCl and the permeating protective additive glycerol. To vary the total initial solute concentration while holding the mass ratio of glycerol to NaCl constant, we had to allow the NaCl tonicity to depart from isotonic. Specifically, human red cells were suspended in solutions with weight ratios of glycerol to NaCl of either 5.42 or 11.26, where the concentrations of NaCl were 0.6, 0.75, 1.0, 2.0, 3.0, or 4.0 times isotonic. Samples were then frozen to various subzero temperatures, which were chosen to produce various molalities of NaCl (0.24-3.30) while holding the fraction of unfrozen water constant, or conversely to produce various unfrozen fractions (0.03-0.5) while holding the molality of salt constant. (Not all combinations of these values were possible). The following general findings emerged: (a) few cells survived the freezing of greater than 90% of the extracellular water regardless of the salt concentration in the residual unfrozen portion. (b) When the fraction of frozen water was less than 75% the majority of the cells survived even when the salt concentration in the unfrozen portion exceeded 2 molal. (c) Salt concentration affected survival significantly only when the frozen fraction lay between 75 and 90%. To find a major effect on survival of the fraction of water that remains unfrozen was unexpected. It may require major modifications in how cryobiologists view solution-effect injury and its prevention.  相似文献   

19.
G Sartor  E Mayer 《Biophysical journal》1994,67(4):1724-1732
Calorimetric studies of the melting patterns of ice in hydrated methemoglobin powders containing between 0.43 and 0.58 (g water)/(g protein), and of their dependence on annealing at subzero temperatures and on isothermal treatment at ambient temperature are reported. Cooling rates were varied between approximately 1500 and 5 K min-1 and heating rate was 30 K min-1. Recrystallization of ice during annealing is observed at T > 228 K. The melting patterns of annealed samples are characteristically different from those of unannealed samples by the shifting of the melting temperature of the recrystallized ice fraction to higher temperatures toward the value of "bulk" ice. The "large" ice crystals formed during recrystallization melt on heating into "large" clusters of water whose redistribution and apparent equilibration is followed as a function of time and/or temperature by comparison with melting endotherms. We have also studied the effect of cooling rate on the melting pattern of ice with a methemoglobin sample containing 0.50 (g water)/(g protein), and we surmise that for this hydration cooling at rates of > or = approximately 150 K min-1 preserves on the whole the distribution of water molecules present at ambient temperature.  相似文献   

20.
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   

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