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 406 毫秒
1.
The influence of dimethyl sulfoxide (DMSO) on the structure and properties of the dipalmitoylphosphatidylcholine membrane was studied at positive temperatures by a combination of X-ray diffraction and small-angle neutron scattering. Penetration of DMSO molecules into the lipid membrane was found to depend on the mole fraction of DMSO in an aqueous solution, X DMSO. At X DMSO > 0.08 the SO group penetrates into the bilayer polar region, thus resulting in structural alterations. At X DMSO > 0.2 defects in the membrane surface are developed.  相似文献   

2.
Small-angle neutron scattering spectra of a polydispersed population of dipalmitoylphosphatidylcholine (DPPC) unilamellar vesicles in heavy water in the presence of dimethyl sulfoxide (DMSO) are analyzed by means of the separated form-factor method. An increase in the mole fraction of DMSO in water from 0 to 15% was shown to lead to an increase in the thickness of the bilayer to the characteristics repeat distances of DPPC multilamellar membranes. This fact is indicative of dehydration of the intermembrane space and a steric contact between adjacent DPPC bilayers at 15% mole fraction of DMSO.  相似文献   

3.
The structural changes in the multilamellar lipid membranes of dipalmitoylphosphatidylcholine (DPPC)/cholesterol and DPPC/ceramide VI binary systems during hydration and dehydration have been studied by neutron diffraction. The effect of cholesterol and ceramide on the kinetics of water exchange in DPPC membranes is characterized. Compared to pure DPPC, membranes of binary systems swell faster during hydration (with a characteristic time of ∼30 min). Both compounds, ceramide VI and cholesterol, similarly affect the hydration of DPPC membranes, increasing the repeat distance due to the bilayer growth. However, in contrast to cholesterol, ceramide significantly reduces the thickness of the membrane water layer. The introduction of cholesterol into a DPPC membrane slows down the change in the parameters of the bilayer internal structure during dehydration. In the DPPC/ceramide VI/cholesterol ternary system (with a molar cholesterol concentration of 40%), cholesterol is partially released from the lamellar membrane structure into the crystalline phase.  相似文献   

4.
Model SC membranes (lipid membranes fabricated from the main components of the lipid matrix of the outermost layer of mammalian skin (stratum corneum (SC)) are used to study the general regularities of the nanostructurization of SC lipid matrix. A model membrane from the outermost layer of skin SC with a composition ceramide 1/ceramide 6/cholesterol/palmitic acid/cholesterol sulfate and a component weight ratio of 30/30/20/15/5 has been investigated by X-ray diffraction. It is shown that the membrane structure at pH = 7.2 consists of two short-periodicity phases with repeat distances d = 47 and 35.7 Å and a long-periodicity phase with d = 127 Å. An increase in pH to 9.0 leads to destruction of the long-periodicity phase, while the short-periodicity phase is retained with d = 48.3 Å.  相似文献   

5.
The influence of the used substrate and the deposition temperature, Td, of the amorphous starting material on the transport properties of solid state crystallized poly-Si is investigated. For poly-Si on a-SiNX:H coated borofloat glass a very small free-carrier concentration that is independent of Td is observed and charge transport is governed by intra-grain scattering mechanisms. For poly-Si on Corning glass the carrier concentration exhibits an inverted U shape as Td increases. These samples exhibit activated behavior of the Hall mobility that is consistent with thermionic carrier emission over potential barriers.  相似文献   

6.
The temperature dependences of birefringence in thin [N(CH3)4]2Zn0.58Cu0.42Cl4 crystals in the thickness range 20 × 10?6d ≤ 500 ? 10?6 m have been investigated. An increase in the temperatures of the parent phase-incommensurate phase and incommensurate phase-ferroelectric phase transitions has been found. The reasons for the shift in the phase transition temperatures with a decrease in the thickness of [N(CH3)4]2Zn0.58Cu0.42Cl4 crystal and the size effect in crystals with an incommensurate superstructure are discussed.  相似文献   

7.
Four liquid crystal compounds of the form 2-sec-butyl-4-[(4-X-phenyl) diazenyl) phenyl-4-(octadecyloxy] benzoate symbolized as I18a, I18b, I18c, and I18d were prepared in which the substituent (X) was taken CH3O-, CH3-, Br-, and -NO2, respectively. Characterization of prepared compounds is done using FT-IR, 1H-NMR, mass spectroscopy, and elemental analysis. Their mesophase was investigated by differential scanning calorimetry. Their antioxidant efficiency for Egyptian lubricating base oil was tested via monitoring the oxidation reaction through the change in total acid number. The obtained results showed that the efficiency of these compounds was ranked as follows I18d > I18c > I18b > I18a.  相似文献   

8.
We report on photocarrier transport of high-growth-rate microcrystalline Si (μc-Si) in conjunction with the lateral size, σL, of crystallites’ conglomerate (grain) determined from the atomic force microscope (AFM) topographic images on the basis of fractal concepts. μc-Si films were prepared using very-high-frequency plasma-enhanced chemical vapor deposition at a high deposition rate of 6.8 ± 0.5 nm/s. μc-Si thicknesses, d, were varied from 0.53 μm to 5.6 μm. With an increase in d, σL increased from 70 nm to 590 nm. At the same time, the ambipolar diffusion lengths, Lamb, of photocarriers, observed using the steady-state photocarrier grating (SSPG) technique, increased from 50 nm to 420 nm. Log–log plots of Lamb versus d and σL versus d were both expressed as a power law with an exponent of 0.9, yielding a simple linear relation between Lamb and σL. Moreover, their ratio, Lamb/σL, was below unity, implying the intra-grain carrier diffusion. From these results, the role of the grain (column) boundaries for photocarrier diffusion in μc-Si is discussed.  相似文献   

9.
The nonvalent interactions in the crystal structures of 16 conformational polymorphs of the C a H b N c O d S e composition (a, b, c, d, and e are stoichiometric indices), including 5-methyl-2-[(2-nitrophenyl)amino]-3-thiophenecarbonitrile (“ROY”) (which has seven different modifications), have been analyzed using the characteristics of molecular Voronoi-Dirichlet polyhedra. It is established that a unique combination of implementable types of intra- and intermolecular nonvalent contacts corresponds to each conformational polymorph.  相似文献   

10.
ESR and optical absorption of Cu2+ were measured in xNa2O(100?x)B2O3 (1 ≤ x ≤ 75), x ZnO(100?x)B2O3 (46 ≤ x ≤ 64) and x Pb(100?x)b2O3 (20 ≤ x ≤ 75) glasses, where x is expressed in mol.%. Spin hamiltonian parameters and ligand field absorption energy changed abruptly in the regions of 15 ≤ x ≤ 23 and 37 ≤ x ≤ 55 in the soda system, while both parameters were hardly dependent upon the glass composition in zinc and lead systems. The magnitude of micro-environmental fluctuation of Cu2+-complexes in the glasses was estimated qualitatively and correlated with the distribution of the strength of π-bonding between cupric ion and oxygen in the glass. Typical network modifiers and intermediates behaved differently, especially in the composition region of invert glass; the large deformation of the coordination sphere of Cu2+ in lead glasses due to the stronger PbO bond resulted in the large distribution of g. The situation was reverse in the case of soda glasses.  相似文献   

11.
The measurements of the Raman spectra on liquid GeXS1?X (0≤X≤0.1) were carried out. The A1 mode of GeS42 (ca. 340 cm?1) was observed in the liquids. The comparison with the spectra of liquid S suggests that Ge reacts on the polymeric species. During the chemical reaction of solid Ge and liquid S below ca. 500°C, the L.T.GeS2 crystal growth was observed. The spectral features of the L.T.GeS2 surprisingly resemble those of GeS glasses. The glass forming region was extended to X = 0.03 by careful rapid quenching. The overall compositional dependence of the Raman spectra of GeXS1?X (0.03≤×≤0.341) was analysed in comparison with the spectra of the L.T. and the H.T. crystal and can be explained by the model that with increase of X(>ca. 0.2) the H.T. GeS2-like spectra emerge abruptly from the L.T.GeS2-like spectra in lower X. Standing on the model, the glassy structures in this system are discussed.  相似文献   

12.
《Journal of Crystal Growth》2006,286(1):102-107
Gallium orthophosphate is a piezoelectric material with an α-quartz structure. In order to manufacture bulk acoustic wave devices (BAW), controlled chemical dissolution is often used to reduce the thickness of the piezoelectric membranes. This paper presents the kinetics of the chemical exchanges, which occur at the solid–liquid interface during the chemical dissolution of GaPO4 in phosphoric acid. Based on chemical composition of phosphoric acid solvent, the pure dissolution rate is determined. A strong anisotropy of chemical reactivity is formed. The dissolution rate is the lowest for the crystallographic z-plane (0 0 0 1) but this orientation is the most sensitive with respect to the proton concentration and the temperature. In accordance with the crystal growth rates, the nucleation at the interface for the (1 0 2 0) plane, named X-plane, is also the most rapidly dissolved. Assuming the activation energies corresponding to dissolution and to nucleation are like standard activation energies, the different values of the standard enthalpy variation are calculated. The most important variation is obtained for the z-plane (ΔrH=−14.3 kJ/mol) and the lowest for the X-plane (ΔrH=−5.4 kJ/mol).  相似文献   

13.
《Journal of Non》2006,352(28-29):3052-3057
Glucofuranose-derivative based gels with various gelator concentration were studied using ultra-small angle X-ray scattering (USAXS) in the scattering vector range 0.0028–0.25 nm−1. A complex method of data processing was applied to extract the structural parameters of the gelator, such as the mass fractal and surface fractal dimensions, dm and ds, respectively, the radius of gyration, Rg, the distance distribution function, p(r), as well as the volume distribution function, D(R), of particles forming aggregates. The results of these analyses show that the aggregate structure changes with the gelator concentration: when it increases, the size of the aggregates decreases, rough aggregates of the fractal type evolve to better developed smooth ones, and their shape changes from disk-like to rod-like. These results suggest that only two different types of aggregates are formed depending on the gelator concentration, and a gradual change between them is observed.  相似文献   

14.
A. Paul  N. Yee 《Journal of Non》1977,24(2):259-276
The vanadium(IV)-vanadium(V) equilibrium in a 37.5BaO, 5.0Al2O3, 57.5B2O3 mol% + X mol% V2O5 (where X = 0.25?32.5) glass system has been studied as functions of temperature, partial pressure of oxygen and total vanadium concentration of the melt. The vanadium(V)/vanadium(IV) ratio in the melt increased with increasing partial pressure of oxygen, lowering temperature of melting, and with increasing total vanadium content of the melt. With X ? 10, the vanadium(V)/vanadium(IV) ratio became almost independent of the total vanadium content of the melt.With this knowledge of oxidation-reduction behaviour, a series of glasses containing 2.8?32.5 mol% V2O5 (at about 4 mol% intervals) and having a constant vanadium(IV)/vanadium(V) ratio (0.17) were prepared. Density, electronic absorption spectrum (both d-d and charge transfer transitions), and ESR of these glasses were measured. Optical and ESR spectra of these glasses indicated the vanadium(IV) to be present as vanadyl ion, VO2+; g| decreased monotonically with increasing vanadium content of these glasses, whereas gβ remained unchanged. The charge transfer transition energy due to vanadium(V) decreased, and the extinction coefficient increased by orders of magnitude with increasing vanadium content of the glass; the most striking changes occurred at X ≈ 10 mol%. DC conductivity of these glasses was measured at different temperatures; a plot of log (?/T) versus 1/T produced straight lines. The slope of these lines remained almost constant (39 ± 1 kcal/mol) for the glasses containing up to about 10 mol% V2O5; with further increase of V2O5 the slope decreased sharply.It has been concluded that the abrupt changes in properties like partial molar volume of V2O5, charge transfer spectrum of vanadium(V), activation energy of polaron hopping — all of which occurred around X ≈ 10 mol% — is due to a major change in the nature of vanadate groups rather than vanadium(IV) in these glasses.  相似文献   

15.
M. Naka 《Journal of Non》1980,41(1):71-77
The crystallization temperature of Fe80?XMXP13C7 and Fe80?XNXP13C7 alloys has been investigated, where M is Zr, Nb, Mo, Ru, Rh or Pd, and N is B, C, Si, P, As, Sn or Sb. Zr and Mo effectively increase TX and Pd decreases it. The increase in crystallization temperature TX with the concentration of elements M, is larger for the second group of periodic elements, except for the case of Pd, compared with the first group of periodic elements. The crystallization temperature is related to the average electron concentration, the atomic size and the atomic cohesive energy of the additive metal.The metalloidal elements N, except for Sn and Sb, increase the TX of Fe80?XNXP13C7 alloys. In particular, the addition of B, C, Si of P definitely increases TX and small amounts of Sn or Sb markedly decrease it. Contrary to the effects of metallic elements, the effects of metalloidal elements do not show the general trend. However, in the same group of the periodic system TX decrease with increasing the atomic size and decreasing cohesive energies for metalloidal elements.  相似文献   

16.
Samples of CZ n-Si〈Zn〉(111) are prepared by high-temperature zinc-diffusion annealing followed by quenching and are studied by X-ray diffraction. The silicon contains an initial phosphorus impurity and zinc-compensating admixture at concentrations N P = 1.5 × 1014 cm?3 and N Zn = 1 × 1014 cm?3; i.e., the relation N P/2 < N Zn < N P is fulfilled. Microdefects are studied by double- and triple-crystal X-ray diffraction in the dispersion free modes (n, ?n) and (n, ?n, +n). The samples are found to contain microdefects with two characteristic sizes (average sizes of about 1 μm and 70 nm). The interplanar distance in the near-surface layer with a thickness of 0.1 μm is smaller than this parameter in the remaining matrix, the difference being equal to d 0 Δd/d 0 ≈ 2 × 10?5. This layer contains mainly vacancy-type microdefects. The angle between the reflecting planes and the local surface relief is Δψ = (7 ± 1) arcmin.  相似文献   

17.
《Journal of Non》2006,352(30-31):3191-3195
Extended X-ray absorption fine structure analyses were carried out on Si1−XGeX films of different thicknesses, prepared by the reactive thermal chemical vapor deposition (CVD) method. From a Rutherford backscattering measurement, the Ge fraction was found to be high near the substrate interface. The Ge coordination ratio, Ge–Ge bond length and Ge–Si bond length decreased with increasing film thickness. The Ge fraction dependences of these parameters were found to be different from the results of previous studies on Si1−XGeX films prepared by molecular beam epitaxy. Our results are considered to be caused by the local structure formation around the Ge atoms during the reactive thermal CVD process.  相似文献   

18.
Crystals of Pb1 ? x BaxSc0.5Nb0.5O3 solid solutions with 0 ≤ x ≤ 0.58 have been grown by the method of mass crystallization from flux. It is established that, unlike the concentration dependence of the corresponding ceramic, the concentration dependence of the temperature T m (the maximum dielectric constant ε in crystals) does not attain saturation. Cooling of crystals with x ≤ 0.04 resulted in a spontaneous transition from the relaxor to the macrodomain ferroelectric state. In crystals with a higher barium content, the relaxor state is “locked in.”  相似文献   

19.
The X-ray crystal structure of a racemic mixture of D- and L-penicillamine has been determined. Crystals are monoclinic,P21/c (No. 14), with cell dimensionsa=11.624(3),b=5.919(1),c=11.482(2) Å,β=114.48(2)°, andZ=2, based on the racemate. The structure was determined by standard methods and refined toR 1=0.0666,R 2=0.0726 for 985 independent reflections. Bond lengths and bond angles do not differ from those in similar structures. Mass spectra and1H and13C NMR spectra are reported ford-penicillamine, and detailed infrared and Raman spectra are reported for solidd-penicillamine hydrochloride,d 5-d-penicillamine hydrochloride,d-penicillamine,d 4-d-penicillamine, anddl-penicillamine. The Raman spectrum ofd-penicillamine in H2O solution as a function of pH is also reported.  相似文献   

20.
The 1∶1 crystal complex of salicylic acid (C7H6O3) and urea (CH4N2O), mp 121° C, is monoclinic, with space groupC2/c (C 2 6 h, No. 15) and unit cell dimensionsa=22.206(3),b=5.108(1),c=17.177(2) Å,β=106.18(1)°.d calc=1.407 g cm?3,d meas=1.41 g cm?3 forZ=8. The structure was determined by direct methods and refined by a full-matrix least-squares procedure to giveR=0.057 andR w =0.050 for 1652 integrated intensities above 2σ(I). The structure contains a strong OH?O hydrogen bond with O?O distance 2.54 Å in which the carboxyl OH group is donor and urea oxygen atom is acceptor. There are two NH?O intermolecular hydrogen bonds with N?O distances of 2.90 and 2.96 Å. Additionally, the salicyclic acid contains an intramolecular OH?O hydrogen bond of 2.56 Å.  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