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 546 毫秒
1.
Cryoprotectors (propylene glycol), ethylene glycol, polyethylene glycol-1500 and dimethyl sulphoxide) are studied for their effect on permeability of liposomes for incorporated molecules of 5,5-dithiobis-2-nitrobenzoic acid (DTNB) under cooling within a temperature range from 0 degree C to -25 degrees C. A similarity is found in the way of ethylene glycol and propylene glycol, dimethyl sulphoxide and polyethylene glycol-1500 effect on the liposome permeability way. Cooling in the presence of ethylene glycol and propylene glycol causes changes in liposome permeability with a local maximum at -18 degrees C. In the medium with 2M NaCl and ethylene glycol, liposomes were resistant to cooling. Dimethyl sulphoxide and polyethylene glycol-1500 induced a two-phase kinetics of changes in liposome permeability, the first phase being within the 0 = -9 degrees C and the second--within -9--25 degrees C temperature ranges. The found differences are supposed to be associated with the effect of the cryoprotective compounds on the lipid crystallization in a lower-temperatures range.  相似文献   

2.
The objectives of this study were to: (a) test the functional activity of Chinchilla lanigera spermatozoa suspended in either glycerol or ethylene glycol, cooled to 4 degrees C, and stored for 24 or 72 h and (b) investigate, after these cooling periods, the effects of incubating sperm at 37 degrees C (for 4 h) upon sperm functional activity. The ejaculate was mixed with the cryoprotectant medium (at 1 M final concentration) and cooled to 4 degrees C. After warming, sperm motility, sperm viability, hypoosmotic swelling test results, and acrosomal integrity were significantly higher for samples containing ethylene glycol than for those in glycerol, stored for 24 or 72 h, and then assayed after 0 or 4 h incubation at 37 degrees C. A significant reduction of sperm motility and viability was detected only when the glycerol cryoprotectant agent was employed, compared to the fresh samples. These results clearly indicate that under our experimental conditions, ethylene glycol is a better protectant for sperm storage than glycerol.  相似文献   

3.
Upon increasing the concentration of ethylene glycol (EG) at 37 degrees C, the twist energy parameter, E(T), which governs the supercoiling free energy, was recently found to undergo a decreasing (or reverse) sigmoidal transition with a midpoint near 20 w/v % EG. In this study, the effects of adding 20 w/v % EG on the torsion elastic constant (alpha) of linear p30delta DNA and on the hydrodynamic radius (R(H)) of a synthetic 24 bp duplex DNA were examined at both 40 and 20 degrees C. The time-resolved fluorescence intensity and fluorescence polarization anisotropy (FPA) of intercalated ethidium were measured in order to assess the effects of 20 w/v % EG on: (1) alpha; (2) R(H); (3) the lifetimes of intercalated and non-intercalated dye; (4) the amplitude of dye wobble in its binding site; and (5) the binding constant for intercalation. The effects of 20 w/v % EG on the circular dichroism (CD) spectrum of the DNA and on the emission spectrum of the free dye were also measured. At 40 degrees C, addition of 20 w/v % EG caused a substantial (1.27- to 1.35-fold) increase in alpha, a significant change in the CD spectrum, and a very small, marginally significant increase in R(H), but little or no change in the amplitude of dye wobble in its binding site or the lifetime of intercalated dye. Together with previously reported measurements of E(T), these results imply that the bending elastic constant of DNA is significantly decreased by 20 w/v % EG at 40 degrees C. At 20 degrees C, addition of 20 w/v % EG caused a marginally significant decrease in alpha and very little change in any other measured properties. Also at 20 degrees C, addition of 30 w/v % betaine caused a marginally significant increase in alpha and significant but modest change in the CD spectrum, but very little change in any other properties.  相似文献   

4.
The aim was to assess the in vitro effect of glycerol, ethylene glycol or acetamide on frozen-thawed ram spermatozoa. Aliquots of each sixteen ejaculates from four rams of the Morada Nova breed were diluted in Tris-egg yolk with glycerol (5%), ethylene glycol (3% or 5%) or acetamide (3% or 5%) and frozen at -196°C. After thawing, progressive sperm motility was greater (P<0.05) in cryopreservation with glycerol 5% and ethylene glycol (3% or 5%) than with acetamide (3% or 5%). Acrosome integrity was greater (P<0.05) with ethylene glycol 5% than acetamide (3% or 5%). The percentage of sperm without oxidative stress was greater (P<0.05) with ethylene glycol (3% or 5%) than with acetamide (3% or 5%). Plasma membrane integrity was greater with glycerol 5% (P<0.05) than with the other cryoprotectants. Thus, it is concluded that glycerol 5% and ethylene glycol 3% or 5% protect ram sperm against the harmful effects of freezing and that glycerol 5% offers greater protection to sperm plasma membrane.  相似文献   

5.
We developed a simple vitrification technique for bovine embryos that could permit direct transfer. Embryos were produced in-vitro by standard procedures. The base medium for cryopreservation was a chemically defined medium similar to SOF + 25 mM Hepes and 0.25% fatty acid free bovine serum albumin (FAF-BSA) (HCDM2). In experiment 1, embryos were first exposed to 3.5M ethylene glycol (V1) for 1, 2 or 3 min at room temperature (20-24 degrees C), and then moved to 7 M ethylene glycol (V2) at 4 or 20-24 degrees C and loaded in 0.25-mL straws. After 45 s in 7 M ethylene glycol, straws were placed in liquid nitrogen. Embryos that were loaded at 20-24 degrees C had higher survival rates than those loaded at 4 degrees C (P<0.05). Exposure for 1 min was best for morulae, while 3 min was best for blastocysts. In experiment 2, blastocysts were handled at 24 degrees C and exposed to two concentrations of ethylene glycol in V1 (3.5 or 5 M) followed by V2 as in experiment 1, two warming temperatures (20 or 37 degrees C) and two post-warming holding times until culture (5 or 15 min). Exposure to 5 M ethylene glycol and warming at 37 degrees C was the optimal combination of procedures, and embryos survived well after 15 min in straws if warmed at 37 degrees C. In experiment 3, ethylene glycol concentration (3, 4 or 5 M) and exposure time (0.5 or 1 min) during two-step addition of cryoprotectant were studied for bovine morulae. In experiment 4, morulae were exposed to V2 for 30 or 45 s in HCDM2 or Vigro holding medium and then held in 22-24 degrees C air or 37 degrees C water post-warming. Experiment 5 was like experiment 4 except blastocysts were used. Overall survival rates of blastocysts in experiment 5 averaged 80% of non-vitrified controls after 48 h culture. The survival rates with in vitro-produced morulae in experiments 1, 3 and 4 were unacceptable. Vitrification solutions based on Vigro tended to result in higher survival than HCDM2 for blastocysts, but not morulae. In experiment 6, the survival rate in vitro of in vivo-produced morulae and blastocysts after two-step vitrification was nearly 100%. Our vitrification technique was very effective for in vitro produced blastocysts, but not for in vitro-produced morulae.  相似文献   

6.
Whole organs, such as kidneys, must be thawed quickly and uniformly to prevent damage during thawing due to excessive heating. Electromagnetic heating with microwaves thaws the kidneys quickly but frequently produces "hot spots" with heat damage. To study heat damage, phantom gelatin kidneys with different dielectric constants and canine kidneys perfused with 12.5% glycerol, ethylene glycol, or dimethyl sulfoxide before freezing were microwave thawed, and the interior temperature was measured by thermography. Phantom kidneys were thawed free standing and canine kidneys were either free standing or packed in a gel mixture. Both phantom and canine kidneys were split symmetrically and separated with a sheet of Styrofoam to facilitate immediate separation and evaluation of the halves after thawing (approximately 3 sec). All the phantoms, regardless of dielectric properties, had areas less than 0 degrees C or greater than 37 degrees C after thawing. The free-standing canine kidneys and the gel-packed ethylene glycol-perfused kidneys had frozen areas (less than 0 degrees C) and hot spots (greater than 37 degrees C). However, glycerol- and dimethyl sulfoxide-perfused kidneys packed in gel before thawing had no areas less than 0 degrees C or greater than 37 degrees C. Altering the geometry from a "kidney shape" to a cylindrical shape with increased volume improved the uniformity of thawing and was more effective than altering the dielectric constant over the range evaluated.  相似文献   

7.
The influence of temperature and water activity (aw) on growth and patulin production by Byssochlamys nivea in apple syrups was determined over a 44-day incubation period. The minimum aw at which the mold was capable of growing was 0.915 and 0.886 at 21 and 30 degrees C, respectively. Growth at 37 degrees C was observed at 0.871 aw. Minimum aw values for patulin production were 0.978, 0.968, and 0.959 at 21, 30 and 37 degrees C, respectively.  相似文献   

8.
The influence of temperature and water activity (aw) on growth and patulin production by Byssochlamys nivea in apple syrups was determined over a 44-day incubation period. The minimum aw at which the mold was capable of growing was 0.915 and 0.886 at 21 and 30 degrees C, respectively. Growth at 37 degrees C was observed at 0.871 aw. Minimum aw values for patulin production were 0.978, 0.968, and 0.959 at 21, 30 and 37 degrees C, respectively.  相似文献   

9.
The supercoiling free energy of pUC19 DNA [2686 base pairs (bp)] was measured in various concentrations of PEG 8000 (polyethylene glycol; molecular weight 8000) by the topoisomer distribution method. The effective twist energy parameter (E(T)) that governs the supercoiling free energy declined linearly by 1.9-fold with increasing w/v % PEG from 0 to 7.5%, which lies below the threshold for intermolecular condensation. In principle, PEG could affect E(T) either via an osmotic exclusion mechanism or by altering the torsion elastic constant, bending rigidity, or self-repulsions of the DNA. Possible alterations of the DNA secondary structure and torsion elastic constant were assessed by CD spectroscopy and time-resolved fluorescence polarization anisotropy of intercalated ethidium. Up to 7.5% PEG, the secondary structure of the DNA remained largely unaltered, as evidenced by (1) the absence of any significant change in the CD spectrum, (2) an extremely small relative decrease (-0.0013) in intrinsic twist, and (3) a negligibly small change in the torsion elastic constant. The observed reduction in E(T) cannot be ascribed primarily to a decrease in torsion elastic constant, and most likely does not stem from a decrease in bending rigidity either. The decrease in medium dielectric constant due to PEG should increase the self-repulsions, and thereby increase E(T), which is opposite to the observed trend. Instead, the observed decline in E(T) is attributed to an osmotic exclusion mechanism. The change in molar volume excluded to the PEG (Delta V(ex)), when the linking difference converts from Delta l = 0 to Delta l = +/-1, was determined from the observed E(T) value and PEG osmotic pressure at each concentration. The experimental Delta V(ex) values agree well with theoretical estimates reckoned for a simple osmotic exclusion model, in which PEG is excluded by hard-core interactions from a concentric cylindrical volume around every duplex segment. The difference in volume excluded to PEG between the Delta l = 0 and the Delta l = +/-1 topoisomers is attributed entirely to the approximately 0.7 additional writhe "crossing" of two duplex strands at roughly 90 degrees, which is known to occur in the latter species. When the separation between the duplex centers at the "crossing" was adjusted so that the theoretical estimate of Delta V(ex) matched the experimental value at each PEG concentration, a value near 5.7 nm was obtained in each case. The invariance and plausible magnitude of this mean separation at the crossing provide strong support for this simple osmotic exclusion model. An alternative model, in which the PEG is excluded from the entire coil envelope of the DNA out to its radius of gyration, perhaps because it decreases the local dielectric constant, was also considered. The estimated difference in excluded volume in that case exceeds the experimental value by a factor of nearly 10(4), and could be ruled out on that basis.  相似文献   

10.
Redox potentials of ferro-ferricyanide and cytochrome c were measured in water/ethylene glycol and water/dimethylsulfoxide (volume ratio from 100/0 to 50/50) between 25 and -25 degrees C. For both systems, the midpoint potential decreases in the presence of organic solvents and increases by cooling. The magnitude of these variations is larger in dimethylsulfoxide than in ethylene glycol; moreover in the same solvent mixture it is larger with ferro-ferricyanide than with cytochrome c, so that the difference between the redox potentials of these two systems can be strongly affected and even reversed. While in pure water (cacodylate buffer pH 7.0, NaCl 0.1 M) they are respectively +388 and +265 mV, in 50% dimethylsulfoxide at 25 degrees C they decrease to +112 and +208 mV. Reduction of cytochrome c by ferro-ferricyanide, in this mixture, is then expected and was indeed observed. On the other hand, as (deltaE/deltaT)T, (E being the redox potential) is higher for ferro-ferricyanide than for cytochrome c, the oxidative power of the former for the latter is expected to increase as temperature decreases. This effect was observed in 50% ethylene glycol at -16 degrees C. Organic solvents and large temperature variations appear then as powerful perturbants of redox reactions. Their effects should be taken into account in studies of redox reactions carried out in cooled hydro-organic media.  相似文献   

11.
We studied the properties of a series of phosphatidylcholine molecules with branched acyl chains. These lipids have previously been shown to have marked stimulatory effects on the side-chain cleavage activity of cytochrome P450SCC (CYP11A1), an enzyme of the inner mitochondrial membrane. The synthetic lipids used were diacyl phosphatidylcholines with the decanoyl, dodecanoyl or tetradecanoyl chain having a hexyl, octyl or decyl straight chain aliphatic branch at the 2-position. All three lipids lowered the bilayer to hexagonal phase transition temperature of dielaidoyl phosphatidylethanolamine, the lipids with longer acyl chains being more effective in this regard. As pure lipids all of the forms were found by X-ray diffraction to be predominantly in the hexagonal phase (HII) over the entire temperature range of 7-75 degrees C. The properties of the HII phase were unusual with regard to the small size of the lattice spacings and the small temperature dependence of the spacings. We used tetradecane to relieve hydrocarbon packing constraints to determine the intrinsic radius of curvature of the lipid monolayer. The elastic bending modulus was measured in the presence of tetradecane by introducing an osmotic gradient across the hexagonal phase cylinders with aqueous solutions of poly(ethylene glycol). The elastic bending modulus was found to be higher than that observed with other lipids and to increase with temperature. Both the small intrinsic radius of curvature and the high elastic bending modulus indicate that the presence of these lipids in bilayer membranes will impose a high degree of negative curvature strain.  相似文献   

12.
Cryomicroscopic observations were made of the volumetric behavior and kinetics of intracellular ice formation (IIF) in Drosophila melanogaster embryos in a modified cell culture medium (BD.20) or BD.20 + 2 M ethylene glycol. After rapid cooling to a given temperature, transient volumetric contraction of the embryos during the isothermal period was quantified by computerized video image analysis. Fitting these data to the numerical solution of the volume flux equation yielded estimates of the hydraulic permeability coefficient (Lp) for individual embryos at various subfreezing temperatures. Lp approximately followed an Arrhenius relation between -2 and -9 degrees C, with a value of 0.168 microns/(min-atm) extrapolated to 0 degrees C and an apparent activation energy delta E of 38.9 kcal/mol. IIF during an isothermal period occurred at random times whose characteristic temperature range and kinetics were affected by the presence of ethylene glycol. A stochastic process model developed to fit these data indicated the influence of both time-dependent and instantaneous components of IIF, presumed to be the result of seeding and heterogeneous nucleation, respectively. The presence of 2 M ethylene glycol depressed the characteristic temperature of instantaneous IIF by about 12 degrees C and reduced the rate constant for time-dependent IIF. Comparison with observed incidences of IIF yielded an estimate of the supercooling tolerance of 3 to 5 degrees C.  相似文献   

13.
AIMS: This study compares the effect of temperature (4-37 degrees C) and water activity (aw: 0.99-0.87) and their interactions on the germination rates, lag times prior to germination and mycelial growth 'in vitro' of Penicillium digitatum, P. italicum and Geotrichum candidum, the main postharvest pathogens affecting citrus fruits. METHODS AND RESULTS: Germination and growth were markedly influenced by temperature and aw. Generally, lag times were longer and germination and growth rates were slower when conditions of temperature and aw were far from optimum. All the studied species were able to germinate over a range of 4-30 degrees C at 0.995 aw, although in non-optimal conditions P. digitatum only reached 40-60% of germinated conidia. At low temperatures, P. italicum germinated and grew faster than P. digitatum and G. candidum, particularly at 0.95 aw. Penicillium italicum was also able to germinate and grow in the driest studied conditions (0.87 aw), while G. candidum did not germinate under 0.95 aw. CONCLUSIONS: Knowledge of the ecological requirements of these fungi is important in order to understand their behaviour in natural situations and to predict fungal spoilage on citrus fruits.  相似文献   

14.
The five-helix bundle lambda6-85* is a fast two-state folder. Several stabilized mutants have been reported to fold kinetically near-downhill or downhill. These mutants undergo a transition to two-state folding kinetics when heated. It has been suggested that this transition is caused by increased hydrophobicity at higher temperature. Here we investigate two histidine-containing mutants of lambda6-85* to see if a weaker hydrophobic core can extend the temperature range of downhill folding. The very stable lambdaHA is the fastest-folding lambda repressor to date (k(f)(-1) approximately k(obs)(-1)=2.3 micros at 44 degrees C). It folds downhill at low temperature, but transits back to two-state folding at its unfolding midpoint. lambdaHG has a weakened hydrophobic core. It is less stable than some slower folding mutants of lambda6-85*, and it has more exposed hydrophobic surface area in the folded state. This mutant nonetheless folds very rapidly, and has the non-exponential folding kinetics of an incipient downhill folder even at the unfolding midpoint (k(m)(-1) approximately 2 micros, k(a)(-1)=15 micros at 56 degrees C). We also compare the thermodynamic melting transition of lambdaHG with the nominal two-state folding mutant lambdaQG, which has a similar melting temperature. Unlike lambdaQG, lambdaHG yields fluorescence wavelength-dependent cooperativities and probe-dependent melting temperatures. This result combined with previous work shows that the energy landscapes of lambda repressor mutants support all standard folding mechanisms.  相似文献   

15.
Factors affecting the survivability of bovine oocytes vitrified in droplets   总被引:8,自引:0,他引:8  
Vitrification of bovine oocytes performed using the traditional, in straw system has not given satisfactory results. Although an alternative approach based on minimizing the volume of the vitrified sample has recently resulted in a much more promising survival rate of vitrified oocytes, we attempted to examine some additional factors influencing the survival and subsequent fertilization and development rates of bovine oocytes subjected to vitrification according to the minimum drop size approach. In total, 748 bovine, in vitro matured oocytes were vitrified using VS14 vitrification solution, containing 5.5-M ethylene glycol and 1.0-M sucrose after different pre-equilibration and equilibration protocols performed at 35 degrees to 37 degrees C. Experiment 1 showed no significant toxic effect during pre-equilibration treatments of oocytes in 2%, 4% or 6% ethylene glycol solutions, except the lower cleavage rate of oocytes exposed to 6% ethylene glycol (77.2% vs. 93.9% in control, P< 0.05). In Experiment 2, 12 to 15 min of pre-equilibration treatments in 0%, 1% or 2% ethylene glycol solutions were tested, followed by 30 or 45 sec of equilibration in VS 14 solution and vitrification in droplets of medium dropped directly into liquid nitrogen. The development rate of vitrified oocytes to the blastocyst stage tended to be higher after 30-sec equilibration treatment (9.5%, 13.9% and 13.8% in groups of oocytes pre-equilibrated in 0%, 1% or 2% ethylene glycol solutions, respectively). Experiment 3 tested pre-equilibration treatments in 0%, 1%, 2%, 3%, 4%, 5% or 6% ethylene glycol solutions, followed by 30-sec equilibration and vitrification in droplets. The highest cleavage, blastocyst and hatched blastocyst rates, which were not significantly different from control, were achieved in a group of oocytes pre-equilibrated in 3% ethylene glycol solution (76%, 30% and 15% vs. 89%, 42% and 21% in control, respectively). A healthy calf was born on Feb 22 1999, after transfer of 4 morula/blastocyst stage embryos developed from oocytes vitrified in droplets after pre-equilibration in 3% ethylene glycol solution. We conclude that gentle pre-equilibration of bovine oocytes in diluted, 3% ethylene glycol solution is an important factor improving the effectiveness of vitrification in droplets of bovine oocytes.  相似文献   

16.
An extramitochondrial acetyl-coenzyme-A hydrolase from rat liver is shown to be a cold-labile oligomeric enzyme that undergoes a reversible conformational transition between a dimeric and a tetrameric form in the presence of adenosine 5'-triphosphate or adenosine 5'-diphosphate at 25-37 degrees C, and between a dimeric and a monomeric form at low temperature. The enzymatically active dimer is fairly stable at 25-37 degrees C, but much less stable at low temperature, dissociating into monomer with no activity. At 37 degrees C and low concentrations of enzyme protein (less than or equal to 14 micrograms/ml), the activity decreased rapidly and only 10% of the initial activity remaining after 60 min. Addition of bovine serum albumin or immunoglobulin G to the medium completely prevented inactivation of the dimeric enzyme at low concentration at 37 degrees C, but had little effect on cold inactivation of the enzyme. Cold inactivation of the dimeric enzyme was partially prevented by the presence of various CoA derivatives. The order of potency was acetyl-CoA (substrate) greater than or equal to butyryl-CoA greater than octanoyl-CoA greater than CoA (product) greater than acetoacetyl-CoA. Another enzyme product, acetate, had little effect on cold inactivation. Polyols, such as sucrose, glycerol, and ethylene glycol, and high concentrations of NaCl, KCl, pyrophosphate and phosphate also greatly prevented cold inactivation. Cold inactivation was scarcely affected by pH within the pH range at which the enzyme was stable at 37 degrees C.  相似文献   

17.
High concentrations of membrane permeable cryoprotectants are necessary to protect human polymorphonuclear leukocytes from osmotic stress injury during freezing, but there are reports that some cryoprotectants are chemically toxic. Cells were exposed to various concentrations of glycerol, dimethyl sulfoxide, or ethylene glycol for 5 min to 2 hr at 37, 22 or 0 degree C, adding or removing the cryoprotectant either slowly or rapidly. Assays included cell number recovery, membrane integrity, phagocytosis, microbicidal ability, and chemotaxis. We conclude that (1) 1 and 2 M concentrations generally are not toxic if they are added and removed slowly at 22 degrees C; (2) addition and removal of glycerol at 0 degree C was injurious even at 1 M; (3) slow addition and removal allowed better recovery than rapid addition or removal; (4) salt concentration in cryoprotectant solutions should be adjusted to isotonic on the basis of moles per liter of solution, rather than moles per kilogram of water; (5) the toxicity reported by other investigators can be largely explained by osmotic stress or dilution shock rather than chemical toxicity; and (6) ethylene glycol is the easiest cryoprotectant to add to and remove from these cells.  相似文献   

18.
It has been shown that aquaporin-3, a water channel, is expressed in mouse embryos. This type of aquaporin transports not only water but also neutral solutes, including cell-permeating cryoprotectants. Therefore, the expression of this channel may have significant influence on the survival of cryopreserved embryos. However, permeability coefficients of aquaporin-3 to cryoprotectants have not been determined except for glycerol. In addition, permeability coefficients under concentration gradients are important for developing and improving cryopreservation protocols. In this study, we examined the permeability of aquaporin-3 to various cryoprotectants using Xenopus oocytes. The permeability of aquaporin-3 to cryoprotectants was measured by the volume change of aquaporin-3 cRNA-injected oocytes in modified Barth's solution containing either 10% glycerol, 8% ethylene glycol, 10% propylene glycol, 1.5 M acetamide, or 9.5% DMSO (1.51-1.83 Osm/kg) at 25 degrees C. Permeability coefficients of aquaporin-3 for ethylene glycol and propylene glycol were 33.50 and 31.45 x 10(-3) cm/min, respectively, which were as high as the value for glycerol (36.13 x 10(-3) cm/min). These values were much higher than those for water-injected control oocytes (0.04-0.11 x 10(-3) cm/min). On the other hand, the coefficients for acetamide and DMSO were not well determined because the volume data were poorly fitted by the two parameter model, possibly because of membrane damage. To avoid this, the permeability for these cryoprotectants was measured under a low concentration gradient by suspending oocytes in aqueous solutions containing low concentrations of acetamide or DMSO dissolved in water (0.20 Osm/kg). The coefficient for acetamide (24.60 x 10(-3) cm/min) was as high as the coefficients for glycerol, ethylene glycol, and propylene glycol, and was significantly higher than the value for control (6.50 x 10(-3) cm/min). The value for DMSO (6.33 x 10(-3) cm/min) was relatively low, although higher than the value for control (0.79 x 10(-3) cm/min). This is the first reported observation of DMSO transport by aquaporin-3.  相似文献   

19.
The lambda(6-85)(*) pseudo-wild type of lambda repressor fragment is a fast two-state folder (k(f) approximately 35 microsec(-1) at 58 degrees C). Previously, highly stable lambda(6-85)(*) mutants with k(f) > 30 microsec(-1) have been engineered to fold nearly or fully downhill. Stabilization of the native state by solvent tuning might also tune lambda(6-85)(*) away from two-state folding. We test this prediction by examining the folding thermodynamics and kinetics of lambda(6-85)(*) in a stabilizing solvent, 45% by weight aqueous ethylene glycol at -28 degrees C. Detection of kinetics by circular dichroism at 222 nm (sensitive to helix content) and small angle X-ray scattering (measuring the radius of gyration) shows that refolding from guanidine hydrochloride denatured conditions exhibits very different time scales for collapse and secondary structure formation: the two processes become decoupled. Collapse remains a low-barrier activated process, while the fastest of several secondary structure formation time scales approaches the downhill folding limit. Two-state folding of lambda(6-85)(*) is not a robust process.  相似文献   

20.
The ability of sublethally heat-stressed aleuriospores of Chrysosporium farinicola to form colonies on yeast extract-glucose agar (YGA) supplemented with sufficient glucose, sorbitol, glycerol, and NaCl to achieve reduced water activity (aw) in the range of 0.88 to 0.95 was determined. The effects of the aw of diluent and incubation temperature during recovery and colony formation were also investigated. Aleuriospores harvested from 14-day-old cultures grown at 25 degrees C were less resistant to heat inactivation compared with aleuriospores from 20-day-cultures. Increased populations of heat-stressed aleuriospores were recovered as the aw of YGA was decreased from 0.95 (glucose and glycerol) and 0.94 (sorbitol) to 0.89 and 0.88, respectively. In NaCl-supplemented YGA, populations recovered at an aw of 0.94 were greatly reduced compared with populations detected at an aw of 0.92; no colonies were formed on NaCl-supplemented YGA at an aw of 0.88. Tolerance to aw values above 0.88 to 0.89 as influenced by solute type was in the order of glucose greater than sorbitol greater than glycerol greater than NaCl. Incubation at 20 degrees generally resulted in an increase in recoverable aleuriospores compared with incubation at 25 degrees C or at 30 degrees C for 14 days followed by 20 degrees C for 10 days. The lethal effect of NaCl on heat-stressed aleuriospores was enhanced at 30 degrees C. The retention of viability of aleuriospores held in sucrose-peptone water diluent (aw, 0.936) for 20 min was essentially the same as that observed when aleuriospores were held in peptone water (aw, 0.997).(ABSTRACT TRUNCATED AT 250 WORDS)  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