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 46 毫秒
1.
Differently bound water molecules of dipalmitoylphosphatidylcholine (DPPC)-H2O system were investigated with differential scanning calorimetry (DSC). According to a method previously reported by us, the ice-melting DSC curves of the DPPC-H2O samples of varying water contents were deconvoluted into multiple components, and the ice-melting enthalpies for the individual deconvoluted components were used to estimate average molar ice-melting enthalpies for freezable interlamellar and bulk waters, respectively. With these average molar ice-melting enthalpies, the numbers of differently bound water molecules of the DPPC-H2O system were calculated at varying water contents and were used to construct a water distribution diagram of this system. Furthermore, to evaluate the reliability of the present DSC deconvolution method, 2H-NMR T1 measurements of DPPC-2H2O system were carried out at 5 °C of the gel phase temperature, and components and fractions for differently bound water (2H2O) molecules were estimated from the analysis of nonexponential magnetization recovery curves.  相似文献   

2.
The two-step crystallization of water in multilamellar vesicles (MLVs) of phosphatidylcholines has been investigated. The main crystallization occurs near -15 degrees C and involves bulk water. Contrary to unilamellar vesicles, a sub-zero phase transition is observed for MLVs at -40 degrees C that corresponds to the crystallization of interstitial water, as proved by Fourier transform infrared absorption and differential scanning calorimetry (DSC) experiments. Furthermore, by means of the DSC method and, more specifically, using the enthalpy change values Delta H(sub) at the sub-zero transition, the number of water molecules per 1,2-dipalmitoylphosphatidylcholine (DPPC) molecule giving rise to this transition has been estimated for different H(2)O/DPPC molar ratios. The curve of the molecular fraction of water molecules involved in the sub-zero transition versus the H(2)O/DPPC molar ratio exhibits a maximum for H(2)O/DPPC equal to 27 (40% in mass of water) and tends towards zero for H(2)O/DPPC ratio values approaching that of the swelling limit of the membrane. A smaller enthalpy value of the sub-zero transition is found for 1-oleoyl-2-palmitoyl-3-phosphatidylcholine (OPPC) than for DPPC. This may be explained by the decrease of interstitial water's quantity when the lipid contains an unsaturated chain. When troxerutin, a hydrophilic drug, is added to the DPPC multilayers, the decrease of Delta H(sub) and melting enthalpy of bulk water is attributed to a decrease of the entropy of the liquid phase owing to the network of water molecules surrounding troxerutin molecules. In all cases, the experiments revealed that the sub-zero transition occurs only in the presence of excess water with respect to the swelling limit of membranes. This evidence could be, at least qualitatively, related to an increase of membrane pressure on interstitial water subsequent to bulk water crystallization.  相似文献   

3.
Ionic and neutral polysaccharides with well-defined structures were chosen to investigate the mechanism of water sorption at different relative humidities. From an experimental point of view, the freezing water was determined by DSC when the total sorbed water was obtained from thermogravimetry. The isotherms of sorption and enthalpies of interaction were determined using the combination of a microbalance and a microcalorimeter. It is shown that freezing water appears for P/P0 > 0.85 especially with the neutral polymers. The differential molar enthalpy of interaction is higher for P/P0 < 0.85 corresponding to the fixation of two water molecules forming double H-bonds; this result is confirmed by molecular modelling; saturation is obtained experimentally for 4 water molecules interacting per glucose unit. On ionic polymers, the water retention increases especially over P/P0 0.8 and the enthalpy of interaction is higher for the first water molecules sorbed. For P/P0 0.8, the numbers of bound water molecules found are 2 per glucopyranosyl unit for neutral polysaccharides, 5 for glucuronan and 9–10 for carboxymethylcellulose (CMC) of D = 2 and hyaluronan (HA)  相似文献   

4.
A measurement of 2H spin-lattice relaxation time, T 1, forD2O was performed with a high resolution liquid NMR apparatus fortwo samples of dimyristoylphosphatidylethanolamine (DMPE)-D2Osystem in a full hydration at varying temperatures of –20, –10, and 5 °C, and both components and compositions of differently boundfreezable water molecules were estimated from a best-fitted curve toexperimental inversion recovery data. A choice of the best-fitted curve wasbased on a distribution of weighted residuals for the experimental data. Asingle component was found for a temperature of –20 °C. At 5 °C, where all the freezable water exists in the liquid state, threecomponents were observed to be characterized by T 1 values ofapproximately 20, 100, and 200 ms, respectively. By comparingcompositions of these individual components with those obtained in ourprevious DSC study, it was revealed that the first and secondarycomponents are members of freezable interlamellar water and the last oneis comparable to bulk water.  相似文献   

5.
Differential scanning calorimetry (DSC) has been used to study the effects of repeated freezing and thawing on dipalmitoylphosphatidylcholine (DPPC) vesicles. Aqueous suspensions of both multilamellar vesicles (MLVs) and large unilamellar vesicles (LUVs) were cycled between -37 and 8 degrees C, and for each thawing event, the enthalpy of ice-melting was measured. In the case of MLVs, the enthalpy increased each time the vesicles were thawed until a steady state was attained. In contrast, the enthalpies measured for LUV suspensions were independent of the number of previous thawing events. It was concluded that MLVs in terms of freezing characteristics contain two pools of water, namely bulk water and interlamellar water. Interlamellar water does not freeze under the conditions employed in the present study, and the MLVs therefore experience freeze-induced dehydration, which is the reason for the observed increase in ice-melting enthalpy. Furthermore, the thermodynamic results suggest that the osmotic stress resulting from the freeze-induced dehydration changes the lamellarity of the MLVs.  相似文献   

6.
The enzyme system composed of human neutrophilic myeloperoxidase (H2O2-oxidoreductase, EC 1.11.1.7), H2O2 and Cl-, at pH 4.5 interacts with egg white lysozyme (EC 3.2.1.17) in several stages. In the first stage, occurring at lysozyme to H2O2 molar ratio of 1:1.4-1.8, the lysozyme loses its enzyme activity but does not yield any derivative distinguishable from the native protein on polyacrylamide gel electrophoresis (PAGE). The second stage of oxidation begins at lysozyme to H2O2 molar ratio above 1:5, producing a change in the lysozyme spectrum at 260-290 nm, and yielding protein derivatives with molecular masses equal to multiples of 14.3 kDa, i.e. the lysozyme molecular mass. This implies that an excessive oxidation of lysozyme by the myeloperoxidase-H2O2-Cl- system produces cross-linking of lysozyme molecules to di-, tri-, tetra-, and pentameric structures. At lysozyme to H2O2 molar ratio exceeding 1:12 a water insoluble white product, which consists of a set of lysozyme cross-linked derivatives, is obtained.  相似文献   

7.
Bound Water in Durum Wheat under Drought Stress   总被引:1,自引:0,他引:1       下载免费PDF全文
To study drought stress effects on bound water, adsorption isotherms and pressure-volume curves were constructed for two durum wheat (Triticum durum Desf.) cultivars: Capeiti 8 (drought tolerant) and Creso (drought sensitive). Plants were grown under well-watered and water-stressed conditions in a controlled environment. Differential enthalpy (ΔH) was calculated through van't Hoff analysis of adsorption isotherms at 5 and 20°C, which allowed us to determine the strength of water binding. ΔH reached the most negative values at approximately 0.06 gram H2O/gram dry weight and then increased rapidly for well-watered plants (until 0.10 gram H2O/gram dry weight) or more slowly for drought-stressed plants (until 0.15-0.20 gram H2O/gram dry weight). Bound water values from pressure-volume curves were greater for water-stressed (0.17 gram H2O/gram dry weight) than for well-watered plants (0.09 gram H2O/gram dry weight). They may be estimates of leaf moisture content where ΔH reaches the less negative values and hence some free water appears. With respect to the well-watered plants, tightly bound water tended to be less bound during drought, and more free water was observed in cv Creso compared to cv Capeiti 8 at moisture contents >0.10 gram H2O/gram dry weight.  相似文献   

8.
In an attempt to understand the initial stage of seed imbibition—the wetting stage—we have examined water binding in dry soybean cotyledon tissue using water sorption isotherm curves. The sorption isotherms show three levels of water affinity: a region of strongly bound water at moisture contents below 8%, a region of weakly bound water at moisture contents between 8 and 24%, and a region of very loosely bound water at contents greater than 24%. The enthalpies of the water binding for the three sectors were −6 to −12.5, about −2.5, and about −0.5 kilocalories per mole water, respectively.

The degree of physiological activity in the tissue reflects the level of water binding. O2 consumption is first detectable in the second region of water affinity (8-24% water), and increases dramatically with increasing water content above about 24%. Damage due to imbibing water is greatest when initial seed moisure contents are in the region of strongest water binding. Damage is lessened and finally absent when seed moisture contents are increased to the second and then to the third level of water affinity.

  相似文献   

9.
The formation of water clusters, polyhydrates of nucleotide bases and their associates during simultaneous condensation of water and base molecules in vacuo onto a surface of a needle emitter cooled to 170 K was studied by field ionization mass spectrometry. It was found that different emitter temperatures are characterized by a specific distribution of intensities of cluster currents, depending on the number of water molecules in clusters. These distributions correlate with structural peculiarities and the relative energetics of formation of water clusters, polyhydrates of nucleotide bases and their associates at low temperature. The features observed in mass spectra for clusters m9Ade (H2O)5, m1Ura (H2O)4 and m9Ade m1Ura (H2O)2 are treated as a result of formation of energetically favorable structures stabilized by H-bonded bridges of water molecules. The relative association constants and formation enthalpies of the noncomplementary pairs Ade Cyt, Gua Ura and the associates which model the aminoacid-base complexes m1Ura Gln and m1.3(2)Thy Gln were determined from the temperature dependencies of the intensities of mass spectra peaks in the range 290-320 K.  相似文献   

10.
The hydration of nucleotide bases of m9Ade(A), m1Ura(U) and a complementary pair A.U was studied by field ionization mass-spectrometry at room and low (170 K) temperatures in vacuum. Enthalpies of A.U-pair formation and its monohydrate A.U(H2O) were measured using temperature dependences of association constants. From the analysis of intensities of mass-spectrum peaks, corresponding to monohydrates U(H2O), A(H2O), A.U(H2O), A.U-pair and initial components A, U, and also measured enthalpies it is supposed that monohydration of bases A and U essentially prevents the formation of the coplanar pair A.U. A qualitative information about the structure and energetics of hydrate clusters A(H2O)n, U(H2O)n and A.U(H2O)n for n = 1 divided by 7 was obtained from low temperature mass-spectra. The observed peculiarities in hydrate structures A(n = 5), U(n = 4), A.U(n = 4) are treated as a consequence of cyclization of water molecules around bases.  相似文献   

11.
The system dioleoylphosphatidylcholine (DOPC)-n-dodecane-2H2O was investigated with different nuclear magnetic resonance (NMR) techniques: (a) a tentative phase diagram was determined by 2H- and 31P-NMR, (b) translational diffusion coefficients were determined for the three components with the pulsed magnetic field gradient NMR technique, and (c) order parameters for perdeuterated n-dodecane were obtained by 2H-NMR. n-Dodecane induces the formation of reversed hexagonal (HII) phases at low and high water concentrations, and cubic phases at low water contents. The translational diffusion coefficients of n-dodecane in a cubic phase with 6 mol water per mol DOPC, and in an HII phase with 48 mol water per mol DOPC, were just approximately 2.5 times lower than in pure dodecane. Perdeuterated dodecane gave large quadrupole splittings in a lamellar phase, much smaller in an HII phase at low water contents, and a narrow single peak in an HII phase at high water contents. This latter observation indicates that a large fraction of the dodecane molecules is located in separate regions between the water cylinders. Our results support the model given by Gruner concerning the aggregation of membrane lipids in the presence of hydrophobic molecules.  相似文献   

12.
Protein-water dynamics in mixtures of water and a globular protein, bovine serum albumin (BSA), was studied over wide ranges of composition, in the form of solutions or hydrated solid pellets, by differential scanning calorimetry (DSC), thermally stimulated depolarization current technique (TSDC) and dielectric relaxation spectroscopy (DRS). Additionally, water equilibrium sorption isotherm (ESI) measurements were performed at room temperature. The crystallization and melting events were studied by DSC and the amount of uncrystallized water was calculated by the enthalpy of melting during heating. The glass transition of the system was detected by DSC for water contents higher than the critical water content corresponding to the formation of the first sorption layer of water molecules directly bound to primary hydration sites, namely 0.073 (grams of water per grams of dry protein), estimated by ESI. A strong plasticization of the T(g) was observed by DSC for hydration levels lower than those necessary for crystallization of water during cooling, i.e. lower than about 0.3 (grams of water per grams of hydrated protein) followed by a stabilization of T(g) at about -80°C for higher water contents. The α relaxation associated with the glass transition was also observed in dielectric measurements. In TSDC a microphase separation could be detected resulting in double T(g) for some hydration levels. A dielectric relaxation of small polar groups of the protein plasticized by water, overlapped by relaxations of uncrystallized water molecules, and a separate relaxation of water in the crystallized water phase (bulk ice crystals) were also recorded.  相似文献   

13.
J A Killian  B de Kruijff 《Biochemistry》1985,24(27):7890-7898
The macroscopic organization, lipid head group conformation, and structural and dynamic properties of 2H2O were investigated in dioleoylphosphatidylcholine (DOPC) model systems of varying gramicidin and 2H2O (or H2O) content by means of small-angle X-ray diffraction and 31P and 2H NMR. At low stages of hydration, N less than 6 (N = 2H2O/DOPC molar ratio), a single lamellar phase is observed in which the gramicidin molecules become preferentially hydrated upon increasing N. For 6 less than N less than 12 phase separation occurs between a gramicidin-poor and a gramicidin-rich lamellar phase. This latter phase is characterized by a smaller repeat distance and decreased DOPC head group order. For N greater than 12, the gramicidin-rich lamellar phase converts to a hexagonal HII phase. Thus, hydration of gramicidin is a prerequisite for HII phase formation in the DOPC/gramicidin system. The HII phase is very rich in gramicidin and 2H2O (gramicidin:DOPC:H2O = 1:1.1:0.9 w/w/w). A model is proposed in which self-assembly of hydrated gramicidin molecules into domains of a specific structure plays a determinant role in the formation of the HII phase by gramicidin.  相似文献   

14.
不同钙-醇溶解体系丝素蛋白的制备及表征研究   总被引:1,自引:0,他引:1  
采用 4种中性盐溶液 Ca(NO3)24H2O 甲醇、Ca(NO3)24H2O 乙醇、CaCl2 甲醇 水和 CaCl2 乙醇 水(摩尔比分别为 1∶2、1∶2、1∶2∶8、1∶2∶8)处理蚕丝纤维,透析后经冷冻干燥制成固体,利用SDS PAGE、电镜扫描和红外光谱对制得的固体进行表征。SDS PAGE结果表明:Ca(NO3)24H2O 醇体系降解丝素蛋白较 CaCl2 醇 水体系降解程度高;电镜扫描的结果表明 Ca(NO3)24H2O 甲醇和 CaCl2 乙醇 水溶解体系处理的丝素蛋白溶解比较完全,Ca(NO3)24H2O 甲醇处理的丝素蛋白冻干后为颗粒状,而 CaCl2 乙醇 水处理的丝素蛋白冻干后为片状。红外光谱的结果表明:4种溶液处理后的丝素蛋白构象均介于 β折叠和无规则卷曲之间,从而为丝素蛋白在药物缓释载体领域的应用提供了一定的理论依据。  相似文献   

15.
Differential scanning calorimetry (DSC) and nuclear magnetic resonance (NMR) spectroscopy are applied to characterize the nonfreezable water molecules in fully hydrated D2O/sphingomyelin at temperatures below 0 degrees C. Upon cooling, DSC thermogram displays two thermal transitions peaked at -11 and -34 degrees C. The high-temperature exothermic transition corresponds to the freezing of the bulk D2O, and the low-temperature transition, which has not previously been reported, can be ascribed to the freezing of the phosphocholine headgroup in the lipid bilayer. The dynamics of nonfreezable water are also studied by 2H NMR T1 (spin-lattice relaxation time) and T2e (spin-spin relaxation time obtained by two pulse echo) measurements at 30.7 MHz and at temperatures down to -110 degrees C. The temperature dependence of the T1 relaxation time is characterized by a distinct minimum value of 2.1 +/- 0.1 ms at -30 degrees C. T2e is discontinuous at temperature around -70 degrees C, indicating another freezing-like event for the bound water at this temperature. Analysis of the relaxation data suggest that nonfreezable water undergoes both fast and slow motions at characteristic NMR time scales. The slow motions are affected when the lipid headgroup freezes.  相似文献   

16.
We report the first calorimetric investigation of steroid diamine binding to a DNA duplex. Absorption spectroscopy, batch calorimetry, and differential scanning calorimetry (DSC) have been used to detect, monitor, and thermodynamically characterize the binding of the steroid diamine, dipyrandium, to poly d(AT). The following thermodynamic data for the binding in 16 mM Na+ at 25 degrees C have been obtained: delta G degree = -6.5 kcal/mol, delta H degree = +4.2 kcal/mol, and delta S = +36 e.u. We interpret the endothermic binding enthalpy in terms of steroid-induced conformational changes in the duplex (e.g. "kinking"). The large positive entropy is interpreted in terms of binding-induced release of bound water and condensed sodium ions. The salt-dependence of the binding constant is interpreted in terms of dipyrandium site-binding involving only one of the two charged ends of the steroid. The optical and DSC curves for the unsaturated steroid-poly d(AT) complexes exhibit biphasic behavior. A comparison of the van't Hoff and the calorimetric transition enthalpies reveals that steroid binding reduces the cooperativity of the transition.  相似文献   

17.
Enthalpies of sublimation, DeltaH degrees (subl) and of solution in water, DeltaH degrees (sol) were determined for a series of crystalline 1,3-dimethyl-uracil derivatives substituted at the C5-ring carbon atom with alkyl groups (-C(n)H(2n+1), n = 2-4) and some of their C(5.6)-cyclooligomethylene analogues (-(CH2)(n)-, n = 3-5). From these data. enthalpies of hydration DeltaH degrees (hydr)= DeltaH degrees (sol) - DeltaH degrees (subl) were calculated and corrected for energies of cavity formation in pure liquid water in order to obtain enthalpies of interaction, DeltaH degrees (int) of the solutes with their hydration shells. The latter are discussed together with the recalculated DeltaH degrees (int) for variously methylated uracils, obtained previously according to a simplified correction procedure, in terms of perturbations in the energy and scheme of hydration of the diketopyrimidine ring brought about by alkyl substitution. It was found that each -CH2-group added with an alkyl substitution contributes favorably about -20 kJ mol(-1) toDeltaH degrees (int).This contribution is partially cancelled by the unfavorable contribution to DeltaH degrees (int) connected with removal of some water molecules bound in the first and subsequent hydration layers by an alkyl substituent. This is particularly evident on substitution at the polar side of the diketopyrimidine ring on which water molecules are expected to be bound specifically.  相似文献   

18.
Dasgupta A  Das D  Das PK 《Biochimie》2005,87(12):7353-1119
The catalytic efficiency of trypsin was estimated in cationic reverse micelles as a function of the concentration of water-pool components and aggregate size to determine their independent influence on enzyme activity. The variation in the aggregate size/water-pool size was achieved by changing both the W0 (mole ratio of water to surfactant) and the headgroup area of surfactant through introduction of hydroxyethyl groups at the polar head. The local molar concentrations of water present inside the water-pool ([H2O]wp) of different cationic reverse micelles across varying W0 was estimated using a modified phenyl cation-trapping protocol. The [H2O]wp in cationic reverse micelles (surfactant/isooctane/n-hexanol/water) increases with W0 and attains the molarity of normal water beyond W0=40 irrespective of the nature of headgroup. Concurrently, the catalytic activity of trypsin compartmentalized within the water-pool increases with the increase in [H2O]wp upto an optimal W0=40 in organized solutions of any surfactant. The aggregate size (determined by static light scattering) also increases expectedly with W0 and noticeably with the area of the surfactant headgroup at similar W0. Since the enzyme activity rises both with the increase in water-pool size and [H2O]wp, trypsin's efficiency was compared with these two parameters across reverse micelles of varying surfactant headgroup size at similar W0 to determine their probable independent influence in regulating the enzyme activity. Noticeably, the efficiency of trypsin rises two to ninefold in spite of the [H2O]wp being distinctly lower in case of hydroxyethyl group substituted surfactants compared to cetyltrimethylammonium bromide w/o microemulsions at similar W0. Thus, the influence of the aggregate size possibly plays an important role alongwith the [H2O]wp in modulating the enzyme activity.  相似文献   

19.
A ferric-EDTA complex, prepared directly from FeCl3 or from an oxidized ferrous salt, reacts with H2O2 to form hydroxyl radicals (.OH), which degrade deoxyribose and benzoate with the release of thiobarbituric acid-reactive material, hydroxylate benzoate to form fluorescent dihydroxy products and react with 5,5-dimethylpyrrolidine N-oxide (DMPO) to form a DMPO-OH adduct. Degradation of deoxyribose and benzoate and the hydroxylation of benzoate are substantially inhibited by superoxide dismutase and .OH-radical scavengers such as formate, thiourea and mannitol. Inhibition by the enzyme superoxide dismutase implies that the reduction of the ferric-EDTA complex for participation in the Fenton reaction is superoxide-(O2.-)-dependent, and not H2O2-dependent as frequently implied. When ferric-bipyridyl complex at a molar ratio of 1:4 is substituted for ferric-EDTA complex (molar ratio 1:1) and the same experiments are conducted, oxidant damage is low and deoxyribose and benzoate degradation were poorly if at all inhibited by superoxide dismutase and .OH-radical scavengers. Benzoate hydroxylation, although weak, was, however, more effectively inhibited by superoxide dismutase and .OH-radical scavengers, implicating some role for .OH. The iron-bipyridyl complex had available iron-binding capacity and therefore would not allow iron to remain bound to buffer or detector molecules. Most .OH radicals produced by the iron-bipyridyl complex and H2O2 are likely to damage the bipyridyl molecules first, with few reacting in free solution with the detector molecules. Deoxyribose and benzoate degradation appeared to be mediated by an oxidant species not typical of .OH, and species such as the ferryl ion-bipyridyl complex may have contributed to the damage observed.  相似文献   

20.
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