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.
Hydrostatic pressure has a vital role in the biological adaptation of the piezophiles, organisms that live under high hydrostatic pressure. However, the mechanisms by which piezophiles are able to adapt their proteins to high hydrostatic pressure is not well understood. One proposed hypothesis is that the volume changes of unfolding (ΔVTot) for proteins from piezophiles is distinct from those of nonpiezophilic organisms. Since ΔVTot defines pressure dependence of stability, we performed a comprehensive computational analysis of this property for proteins from piezophilic and nonpiezophilic organisms. In addition, we experimentally measured the ΔVTot of acylphosphatases and thioredoxins belonging to piezophilic and nonpiezophilic organisms. Based on this analysis we concluded that there is no difference in ΔVTot for proteins from piezophilic and nonpiezophilic organisms. Finally, we put forward the hypothesis that increased concentrations of osmolytes can provide a systemic increase in pressure stability of proteins from piezophilic organisms and provide experimental thermodynamic evidence in support of this hypothesis.  相似文献   

2.
Dilatometric measurements of the volume changes accompanying the binding reactions of azide ion to human adult and pigeon methemoglobins as a function pH at 25°C demonstrate pH values of maximum volume change (pH ΔVmax) which are different for the different hemoglobins. pHΔVmax occurs at pH 6.7 for human methemoglobin A and at pH 7.7 for pigeon methemoglobin. The pHΔVmax occurs near the characteristic pH (pHch) of maximum enthalpy of the same binding reaction. It is shown that the large pH variation in ΔV can arise if the configuration of charged groups on the surface of the molecule is different in methemoglobin and methemoglobin complex. When such a difference in configuration exists the addition of the same number of protons to methemoglobin and methemoglobin complex will give rise to different changes in the partial molar volume of the two species.  相似文献   

3.
DNA hairpins consist of two distinct structural domains: a double stranded stem and a single-stranded loop that connect the two strands of the stem. Previous studies of short DNA hairpins have revealed that loop and stem sequences can significantly affect the thermodynamic stability of short DNA hairpins. In this work we present the effect of hydrostatic pressure on the helix-coil transition temperature (TM) for 11 16-base, hairpin-forming DNA oligonucleotides. All of the samples form a hairpin with a 6-base pair stem and a four-base loop. In addition, the four base pairs at the end of the stem distal from the loop are the same for every molecule. We have varied loop sequence and identity of the two duplex base pairs adjacent to the loop. Using the change in UV absorption to monitor the conformational state of the oligonucleotide the hairpin-coil transition temperature of these molecules was studied as a function of sodium ion concentration and pressure. From these data we calculated the volume change accompanying the transition. Model-dependent (van't Hoff) transition parameters such as ΔHvH and transition volume (ΔV) were estimated from the analysis of conformational transitions. Experiments revealed that the ΔV for denaturation of these molecules range from − 2.35 to + 6.74 cm3 mol−1. The expansibility (ΔΔV/ΔT) and the pressure dependence of cation release are also presented. The difference in the volume change for this transition is related to the differences in the hydration of these molecules.  相似文献   

4.
The effect of temperature (20-70 °C) on the gelatinization and retrogradation of potato starch-water mixtures (10-70%, w/w) treated with high hydrostatic pressure (HHP) (400-1000 MPa) was investigated. Gelatinization enthalpy change (ΔHgel) and re-gelatinization enthalpy change of retrograded crystalline part (ΔHretro) of the HHP-treated starch were evaluated using differential scanning calorimetry. The value of ΔHgel of 10-20% (w/w) mixtures decreased with increased pressure and temperature, while ΔHgel of 30-50% (w/w) mixtures decreased to certain values with increased pressure and the values depended on treatment temperature. With higher temperature and pressure conditions, ΔHgel of 10-40% (w/w) mixtures reached zero, but ΔHgel of 50-70% (w/w) mixtures did not. Retrogradation was observed with HHP-treated 20-60% (w/w) mixtures and the value of ΔHretro depended on the starch content, pressure, and temperature. The value of ΔHretro trended to increase with increase in starch content. In addition, retrogradation was promoted by HHP treatment at low temperature. Gelatinizaiton and retrogradation behaviors of HHP-treated (400-1000 MPa) potato starch-water mixtures (10-70%, w/w) at 20-70 °C were summerized in a series of state diagrams.  相似文献   

5.
High hydrostatic pressure inhibits growth in most organisms; this may be explained by a deactivation of enzymes involved in essential metabolic pathways. In order to check this hypothesis the enzymic activity of rabbit muscle lactic dehydrogenase and yeast glyceraldehyde-3-phosphate dehydrogenase was investigated in the presence of the coenzyme and excess of substrate at pressures up to 2 kbar.Kinetic analysis of an initial phase of pressure induced activation and of a second phase of reversible deactivation shows that the two enzymes respond to high pressures in different ways leading to a volume of activation of ΔV3(LDH) = 0 ± 1 cm3 mol−1 and ΔV3(GAPDH) = 60 ± 4 cm3 mol−1, respectively. Comparing the lower limits of pressure deaclivation, LDH is found to be more stable towards pressure than GAPDH. At p ≈ 2 kbar total deactivation of both enzymes is observed. A concentration dependent lag of GAPDH reactivation proves dissociation to participate in the process of deactivation, while the effects for LDH are explicable on the basis of reversible denaturation alone.  相似文献   

6.
The Pressure Dependence of the Helix-Coil Transition Temperature (Tm) of Poly[d(G-C)] was studied as a function of sodium ion concentration in phosphate buffer. The molar volume change of the transition (ΔV) was calculated using the Clapeyron equation and calorimetrically determined enthalpies. The ΔV of the transition increased from +4.80 (±0.56) to +6.03 (±0.76) mL mol?1 as the sodium ion concentration changed from 0.052 to 1.0M. The van't Hoff enthalpy of the transition calculated from the half-width of the differentiated transition displayed negligible pressure dependence: however, the value of this parameter decreased with increasing sodium ion concentration, indicating a decrease in the size of the cooperative unit. The volume change of the transition exhibits the largest magnitude of any double-stranded DNA polymer measured using this technique. For poly[d(G-C)] the magnitude of the change in ΔV with sodium ion concentration (0.94 ± 0.05 mL mol?1) is approximately one-half that observed for either poly[d(A-T)] or poly (dA)·poly(dT). The ΔV values are interpreted as arising from changes in the hydration of the polymer due to the release of counterions and changes in the stacking of the bases of the coil form. As a consequence of solvent electrostriction, the release of counterions makes a net negative contribution to the total ΔV, implying that disruption of the slacking interactions contributes a positive volume change to the total ΔV. The larger magnitude of the ΔV compared with that of other double-stranded polymers may be due in part to the high helix-coil transition temperature of poly[d(G-C)], which will attenuate the contribution of electrostriction to the total volume change. The data in addition show that in the absence of other cellular components, the covalent structure of DNA is stabile under conditions of temperature and pressure more extreme than those experienced by any known organism. © 1995 John Wiley & Sons, Inc.  相似文献   

7.
Lactic dehydrogenase from pig heart can be reversibly dissociated at hydrostatic pressures above 1000 bar. The breakdown of the native quaternary structure occurs at lower pressures compared to the isoenzyme from pig, skeletal muscle. As shown by hybridization experiments of the two isoenzymes the final product of dissociation is the homogeneous monomer. Fluorescence emission spectra of the monomeric enzyme at elevated pressure are characterized by a decrease in fluorescence intensity without any red shift, indicating that no significant unfolding occurs upon high-pressure dissociation. The spectral changes are comparable to those observed after acid dissociation. The amount and rate of deactivation depend on pressure and on the conditions of the solvent. The presence of various anions (Cl?, SO2?4. HPO42?) has no effect on the stability of ihe enzyme towards pressure. High-pressure denaturation (as monitored by intrinsic protein fluorescence), and deactivalion (measured immediately after decompression) run parallel; the pressure dependence of their first-order rate constants is characterized by an activation volume ΔVDc = ?140 = 10 cm3/mol. As taken from the yield of reconstitution, dissociation, denaturation and deactivation are found to be fully reversible provided the pressure does not exceed a limiting value (p = 1000 bar in Tris. pH 7.6: 24 h incubation at 20°C). After extended incubation beyond the limiting, pressure of 1000 bar. “irreversible high-pressure denaturation” occurs which is accompanied bv partial aggregation after decompression. The coenzyme, NAD+ stabilizes the native tetramer shifting the dissociation equilibrium to higher pressures. The overall dissociation-association reaction can be quantitatively described by a consecutive dissociation/unfolding mechanism N?4 M'?4 M (where N is the native tetramer. and M' and M two different conformations of the monomer). The reaction volume of the dissociation reaction N?4 M' is found to be ΔVDiss = ?360 = 30 cm3/mol: as indicated by the pressure dependence of the yield of reconstitution, the reaction volume of the equilibrium M'?MXXX is also negative.  相似文献   

8.
Okumura H 《Proteins》2012,80(10):2397-2416
A multibaric‐multithermal molecular dynamics (MD) simulation of a 10‐residue protein, chignolin, was performed. All‐atom model with the Amber parm99SB force field was used for the protein and the TIP3P model was used for the explicit water molecules. This MD simulation covered wide ranges of temperature between 260 and 560 K and pressure between 0.1 and 600 MPa and sampled many conformations without getting trapped in local‐minimum free‐energy states. Folding events to the native β‐hairpin structure occurred five times and unfolding events were observed four times. As the temperature and/or pressure increases, fraction of folded chignolin decreases. The partial molar enthalpy change ΔH and partial molar volume change ΔV of unfolding were calculated as ΔH = 24.1 ± 4.9 kJ/mol and ΔV = ?5.6 ± 1.5 cm3/mol, respectively. These values agree well with recent experimental results. Illustrating typical local‐minimum free‐energy conformations, folding and unfolding pathways were revealed. When chignolin unfolds from the β‐hairpin structure, only the C terminus or both C and N termini open first. It may undergo an α‐helix or 310‐helix structure and finally unfolds to the extended structure. Difference of the mechanism between temperature denaturation and pressure denaturation is also discussed. Temperature denaturation is caused by making the protein transferred to a higher entropy state and making it move around more with larger space. The reason for pressure denaturation is that water molecules approach the hydrophobic residues, which are not well hydrated at the folded state, and some hydrophobic contacts are broken. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

9.
S Makino  H Noguchi 《Biopolymers》1971,10(7):1253-1260
The measurements were made for the volume and the sound velocity changes (ΔV and ΔU) on titrating the sodium salt of poly (S-carboxymethyl L -cysteine) with dilute HCl. For the reaction, ? COO? + H+ → ? COOH, ΔV per mole of H+ bound was + 12. 7 ml and +11. 4 ml in salt-free and 0. 2 M NaCl solutions, respectively. Corresponding ΔU was about ?13 cm/sec in salt-free polymer solution where 11.5 mM carboxylate ion reacts with equimolar hydrogen ion. ΔV associated with the coil-to-β transition was found to be +2. 35 ml in H2O and +1. 90 ml in 0. 2 M NaCl per mole of amino acid residue, respectively. These values are larger than those obtained for the coil-to-helix transition of poly (L -glutamic acid). ΔU for the transition was about ?30 cm/sec in salt-free solution of polymer concentration 0.0115 mole/liter. Possible sources of ΔV and ΔU for reaction; coil → β, are (1) the formation of void volume and (2) the changes in the extent of solvation in amide linkage and in side chain.  相似文献   

10.
The pressure dependence (10–4000 bar) of the kinetics of the ammoniation of[Co(NH3)5X](ClO4)2 (X = N3, Cl) and the isomerization of [Co(NH3)5(ONO)](ClO4)2 in liquid ammonia is reported. The conjugate-base mechanism is operative for these complexes over the entire pressure range used. Activation and thermodynamic parameters were obtained for each of the two steps of the mechanism for [Co(NH3)5(N3)](ClO4)2 at 20 bar. Values for the overall activation volume extrapolated to zero pressure are ΔV3(0) = ?12 (11.35 °C, ONO); ?20 (24.45 °C, N3) and ?30 (0.50 °C, Cl) cm3 mol?1. Application of El'yanov and Hamann's empirical relation for the pressure dependence of the ionization of weak acids separates the contributions of the pre-equilibrium (ΔVCB0) and the elimination or isomerization reaction (ΔV23) (at zero pressure). The values obtained for [Co(NH3)5X](ClO4)2 are (givens as X; ΔVCB0 and ΔV23 in cm3 mol?1; T in °C): (ONO; ?16 and ?15; 11.35), (N3; ?22 and 1;24.45), (Cl; ?22 and 3;0.50). These values fit in the accepted picture of volume effects in cobalt(III) ammine kinetics.  相似文献   

11.
Upon activation, an internal hydrostatic pressure develops within the Fundulus egg, and compresses the egg proper to a reduced volume. When the perivitelline pressure is abolished by a highly hypertonic sucrose solution, the egg volume increases. As sucrose penetrates the chorion, the volume again decreases. The relation between P and V in these conditions is inverse, and approximates a rectangular hyperbola. The limiting factor causing most of the deviation is shown to be the incompressible fraction. It is concluded that the volume of the egg proper is controlled by the perivitelline pressure, and that the effect of hypertonic sucrose solution is exerted by lowering the pressure and thereby increasing membrane permeability non-specifically. It is also shown that some permanent alterations occur within the plasma membrane during activation that reduce the permeance, and thereby, increase the incompressible fraction.  相似文献   

12.
An ultrafiltration technique was used to study the temperature coefficient of the association constant K for 1:1 binding of proteoglycan to a hyaluronate oligosaccharide fraction containing an average of about 16 monosaccharide units. The proteoglycan was concentrated during the filtration experiment in order to provide minimal disturbance of the equilibrium in the retained solution. Analytical results calculated from assay of 3H-labeled hyaluronate in the filtrate fractions were extrapolated back to initial equilibrium cell conditions. At 10 °C values of K obtained in this way from ultrafiltration agreed within experimental error with those from equilibrium dialysis. Apparent K values obtained with both techniques tended to decrease somewhat with increasing proteoglycan concentration, due probably in part to excluded volume effects. Values of K obtained by ultrafiltration over the temperature range 5 to 40 °C were used to estimate the enthalpy of binding ΔH° as ?17.5 (±1.5) kcal mol?1 and the entropy of binding ΔS° as ?50 (±5) cal K?1 mol?1 (based on a 1 μm standard state). The dilute solution value of K at 37 °C is sufficiently large to suggest that most of the proteoglycan monomers having a binding site are complexed under tissue conditions.  相似文献   

13.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

14.
The effect of hydrostatic pressure on the helix-coil transition temperature (Tm) was measured for the DNA oligomers (dA)n(dT)n, where n = 11, 15, and 19, in 50 mM NaCl. The data were analyzed in light of previously published data for the polymer, poly(dA)·poly(dT) under the same conditions. As has been observed for DNA polymers, increasing the hydrostatic pressure led to an increase in the Tm of the oligomers; however, the effect of pressure diminished with decreasing chain length. The value of dTm/dP decreased linearly with the inverse of the chain length varying from 3.15 × 10−2°C MPa−1 for the polymer to 0.7 × 10−2°C MPa−1 for the 11-mer. The two-state or van't Hoff enthalpy (ΔHvH) of the helix-coil transition was obtained by analysis of the half-width of the thermal transition. As expected, ΔHvH decreases with decreasing chain length. In contrast to the behavior of the polymer, poly(dA)·poly(dT), and (dA)19(dT)19, the ΔHvH of the two shorter duplex oligonucleotides displayed a small pressure dependence dΔHvH/dP≃−0.4 kJ MPa−1 in both cases. The changes observed in the Tm and ΔHvH were not sufficient to explain the magnitude of the chain-length dependence of the pressure effect. To interpret the large chain-length dependence of dTm/dP, we propose that the terminal base pairs contribute a negative volume change to the helix-coil transition. Base pairs distant from the ends exhibit behavior characterized by the polymer where end effects are assumed to be negligible, i.e., a positive volume change for the helix-coil transition. The negative volume change of separating terminal bases may originate from the imperfect interactions these base pairs form with water due to the existence of several energetically equivalent conformations. This is reminiscent of one of the mechanisms proposed to be important in the pressure-induced dissociation of multimeric proteins into their constituent subunits. © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Lactic dehydrogenase (LDH) from pig heart and pig skeletal muscle can be reversibly dissociated into monomers at high hydrostatic pressure. The reaction can be quantitatively filled by a reversible consecutive dissociation-unfolding mechanism according to Na = 4M ? 4M* (where N is the native letramer, and M and M* two different conformations of the monomer) (K. Müller, et al., Biophys. Chem. 14 (1981) 101). At P ? 1 kbar, the pressure deactivalion of both isoenzymes (H4 and M4) is described by the two-state equilibrium N ? 4M. From the respective equilibrium constant and the temperature and pressure dependence of the change in free energy, the thermodynamic parameters of the dissociation/deactivation may be determined, e.g., for LDH-M4: ΔgDiss = 110 kJmol, ΔSDiss = ?860 J/K per mol, ΔHDiss = ?124 kJmol (enzyme concentration 10 μgml, in Tris-HCl buffer, pH 7.6, I = 0.16 M, 293 K, 0.8 kbar); the dissociation volume is found to be ΔVDiss = ?420 mlmol (0.7 < p < 0.9 kbar). Measurements using 8-anilino-1-naphlhalenesulfonic acid (ANS) as extrinsic fluorophore demonstrate that the occurrence of hydrophobic surface area upon dissociation parallels the decrease in reactivation yield after pressurizarion beyond 1 kbar. Within the range of reversible deactivation (p < 1 kbar) no increase in ANS fluorescence is detectable, thus indicating compensatory effects in the process of subunit dissociation. 2H2O is found to stabilize the enzyme towards pressure dissociation, in accordance with the involvement of hydrophobic interactions in the subunit contact of both isoenzymes of LDH.  相似文献   

16.
Anoxybacillus beppuensis TSSC-1 (GenBank Number, EU710556), a thermophilic bacterium isolated from a hot spring reservoir, was found to optimally secrete a monomeric α-amylase at 55 °C and pH 7. The enzyme was purified to homogeneity by a single-step purification on phenyl sepharose 6FF, achieving a 58% yield, 10,000 U/mg specific activity and 19.5 fold purification. The molecular weight, Km and Vmax were 43 kD, 0.5 mg ml?1 and 3571.42 μmol ml?1 m?1, respectively. The enzymatic catalysis of soluble starch was optimum at 80 °C and pH 7. The thermodynamic parameters, Kd, t1/2, ΔH*, ΔS*, E and ΔG*, were consistent. The very compact structure of the enzyme and the transitional enzyme–substrate complex resisted denaturation at extreme temperatures and alkaline pH. The Kd and t1/2 measurements were consistent with the high thermostability and pH tolerance observed. The structural stability of the enzyme was also reflected by the values of ΔH*, ΔS*, E and ΔG*. While the enzyme did not exhibit metal ion dependency, it was resistant to chemical denaturation. The broad thermo- and pH-tolerance of this enzyme suggests potential commercial opportunities.  相似文献   

17.
1. The effects of lipid-protein interactions on carp brain and liver mitochondrial MAO with respect to substrate and inhibitor preference, thermostability and Arrhenius parameters were studied and compared.2. Treatment with phospholipase A2, C or D decreased MAO activities towards 5-hydroxytryptamine (5-HT), β-phenylethylamine and tyramine similarly, accompanied by great changes in their apparent affinities for MAO, but not by changes in Vmax values.3. Minimum phospholipid binding to mitochondria might be essential for enzyme activity.4. Among these activities, 5-HT deamination was the most sensitive to the changes in mitochondrial phospholipids and bulk lipid phase transition (fluidity).5. Sensitivity of MAO to clorgyline or l-deprenyl was not affected by these phospholipase treatments.6. Of the phospholipids tested, only phosphatidylinositol significantly activated MAO activity towards 5-HT in both intact and phospholipase-treated mitochondria.  相似文献   

18.
Lipases are important to high value product synthesis, modification, and enhancement. However, they are often unstable above 40 °C. While most current applications of high hydrostatic pressure (HHP) are for inactivating deleterious enzymes, there is evidence that HHP can stabilize and increase activity of some enzymes. This study examines the apparent kinetics of immobilized lipase-catalyzed synthesis of isoamyl acetate at HHP in hexane. HHP reduced thermal inactivation of lipase by up to 152% after 4 h at 80 °C and 400 MPa when compared to incubations at low pressure. No significant differences were found in activation energy (Ea) at different pressures, irrespectively of the pressurization and heating sequence, and were between 35.7 ± 3.5 and 47.8 ± 8.2 kJ mol?1, depending on the method. In all methods utilized, activity at 63.5 and 80 °C at 400 MPa was greater (from about 20 to 96% increase) than at low pressure. Activity increased by 110% at low pressure versus a 239% increase at 350 MPa when the temperature was increased from 40 to 80 °C. Increasing pressure up to 350 MPa increased lipase activity while pressures greater than 350 MPa maintained or decreased lipase activity. Activation volume (ΔV) appeared negative between ambient pressure and 200 MPa in contrast to a positive ΔV between 300 and 600 MPa. Apparent ΔV was 14.3 ± 1.7 or 15.2 ± 2.2 cm3 mol?1 at 40 or 80 °C, respectively, between 300 and 500 MPa.  相似文献   

19.
Among food-borne pathogens, some strains could be resistant to hydrostatic pressure treatment. This information is necessary to establish processing parameters to ensure safety of pressure-pasteurized foods (N. Kalchayanand, A. Sikes, C. P. Dunne, and B. Ray, J. Food Prot. 61:425–431, 1998). We studied variation in pressure resistance among strains of Listeria monocytogenes, Staphylococcus aureus, Escherichia coli O157:H7, and Salmonella species at two temperatures of pressurization. Early-stationary-phase cells in 1% peptone solution were pressurized at 345 MPa either for 5 min at 25°C or for 5, 10, or 15 min at 50°C. The viability loss (in log cycles) following pressurization at 25°C ranged from 0.9 to 3.5 among nine L. monocytogenes strains, 0.7 to 7.8 among seven S. aureus strains, 2.8 to 5.6 among six E. coli O157:H7 strains, and 5.5 to 8.3 among six Salmonella strains. The results show that at 25°C some strains of each species are more resistant to pressure than the others. However, when one resistant and one sensitive strain from each species were pressurized at 345 MPa and 50°C, the population of all except the resistant S. aureus strain was reduced by more than 8 log cycles within 5 min. Viability loss of the resistant S. aureus strain was 6.3 log cycles even after 15 min of pressurization. This shows that strains of food-borne pathogens differ in resistance to hydrostatic pressure (345 MPa) at 25°C, but this difference is greatly reduced at 50°C. Pressurization at 50°C, in place of 25°C, will ensure greater safety of foods.  相似文献   

20.
The effect of general anaesthetics, alcohols and hydrostatic pressure on the thermal transition in dipalmitoyl phosphatidylcholine multilayer liposomes has been measured using dilatometry. The volume increasse at the transition (ΔVt) is 0.0350 ± 0.0003 ml/g. the transition temperature (Tt) 41.84 ± 0.09°C and the width of the transition 1.025 ± 0.18°C. ΔH calculated by the Clapeyron-Clausius equation is 8.4 kcal/mol. The n-alcohols C3C5 reduced the transition temperature without affecting the transition width which was however, increased by n-hexanol. Trichloroethylene, the fluorescent probe N-phenyl-1-naphthyl-amine, and methoxyflurane all increased the transition width (reduced the cooperativity of the transition) with a simultaneous depression of Tt. Methoxyflurane caused a two-stage transition expansion. Diethyl ether's effect has similarities with both the C3 and C6 alcohols. Generally ΔVt was unaffected by the agents.Pressure increased Tt by 0.0238°C/atm linearly over the range 1–300 atm in both treated and untreated liposomes, and therefore cannot be said to antagonize anaesthetics. In both treated and untreated liposomes ΔVt and the width of the transition were unaffected by pressure. Pressure thus reverses the effects of anaesthetics on Tt but not their spread of the transition width.  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