首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The apparent molar volumes, V of diglycine, triglycine and glycyl-l-leucine have been determined in water and in aqueous sodium acetate (0.5, 1.0, 2.0, and 4.0 mB) and magnesium acetate (0.5, 1.0, 1.5, and 2.0 mB) solutions at 298.15 K by the measurement of densities using vibrating-tube digital densimeter. The partial molar volumes, V2,m0 obtained from V have been used to calculate the partial molar volumes of transfer, ΔtrV2,m0 for these peptides from water to aqueous solutions of sodium acetate (SA) and magnesium acetate (MA) solutions. The hydration numbers, nH and volumetric interaction coefficients have also been calculated. The ΔtrV2,m0 data suggest that ion-charged/or peptide group interactions of peptides are stronger with MA in comparison to SA.  相似文献   

2.
The ultrasonic speeds, u and viscosities, η of binary mixtures of formamide (FA) with ethanol, 1-propanol, 1,2-ethanediol, and 1,2-propanediol, including those of pure liquids, over the entire composition range were measured at 293.15, 298.15, 303.15, 308.15, 313.15, and 318.15 K. From the experimental values of u and η, the deviations in isentropic compressibility, Δks, in ultrasonic speed, Δu, and in viscosity, Δη were calculated. The variation of these parameters with composition and temperature of the mixtures are discussed in terms of molecular interaction in these mixtures. The observed trends in Δks values indicate the presence of specific interactions between FA and alkanol molecules. The Δks values follow the order: ethanol < 1-propanol < 1,2-propanediol < 1,2-ethanediol. It is observed that the Δks values depend upon the number of hydroxyl groups and alkyl chain length in these alkanol molecules. Furthermore, the free energies, ΔG, enthalpies, ΔH and entropies, ΔS of activation of viscous flow have also been obtained by using Eyring viscosity equation and their dependence on composition of the mixtures have been discussed.  相似文献   

3.
Conductivities of some tetraalkylammonium halides, viz. tetramethylammonium iodide (Me4NI), tetraethylammonium bromide (Et4NBr), tetraethylammonium iodide (Et4NI), tetra-n-propylammonium bromide (Pr4NBr), tetra-n-butylammonium bromide (Bu4NBr), tetra-n-butylammonium iodide (Bu4NI) and tetra-n-heptylammonium bromide (Hp4NBr) were measured at 298.15 K in 1,3-dioxolane which has a low permittivity (ε = 7.13). A minima in the conductometric curves (molar conductance, Λ vs. square root of concentration, √c) was observed for concentrations which were dependant upon both the salt and the solvent. The observed molar conductivities were explained by the formation of ion-pairs (M+ + X ↔ MX, KP) and triple-ions (2M+ + X ↔ M2X+; M+ + 2X ↔ MX2, KT). A linear relationship between the triple-ion formation constants [log (KT/KP)] and the salt concentrations at the minimum conductivity (log Cmin) was given for all the salts in 1,3-dioxolane. The formation of triple-ions might be attributed to the ion sizes in solutions in which Coulombic interactions and non-Coulombic interactions act as the main forces between the ions (R4N+…..X).  相似文献   

4.
Densities (ρ) of the binary mixtures of acetic acid with methanol, ethanol, propan-1-ol, propan-2-ol, butan-1-ol and butan-2-ol have been measured at temperature 298.15 K and ambient pressure (815 hPa) as a function of composition using an Anton Paar model DMA 4500 oscillating densimeter. The excess molar volume (VmE), partial molar volume (i) and excess partial molar volumes (iE) of the binary mixtures were calculated from the density data. The excess molar volumes were correlated with the Redlich−Kister equation. The excess molar volumes are negative for methanol, ethanol, propan-1-ol, propan-2-ol and butan-1-ol. They are positive for butan-2-ol and an inversion of sign in VmE is observed for butan-2-ol around 0.9 mol fraction of acetic acid.

The results obtained in this work were interpreted in terms of intermolecular interaction between like and unlike molecules, difference in size and shape of unlike molecules and the steric hindrance caused by increased methylation.  相似文献   


5.
We have measured the conductivity σ of TlX(X=Cl, Br, I) compounds up to 5.3 GPa and between 300–823 K. The σT dependence for all compounds can be divided into three distinct regions: (i) low temperature (LT), <400 K, with unusual negative σT dependence, (ii) intermediate temperature (IT), 400<650 K, with positive σT dependence and (iii) high temperature (HT), T>650 K, with positive σT dependence. The σT isobars were used to construct the TP solid phase diagram for each compound. The LT region data indicate a new meta-stable phase in the 1.0–3.5 GPa range. The LT→IT transition is characterized by an inverse σT dependence followed by normal Arrhenius behavior up to and including the HT region. The extrapolation to 1 atm of the P-dependent boundary between IT and HT regions above 3 GPa for each compound in the PT plot yields a value close to its respective normal (1 atm) Tmelt suggesting a solid order–disorder transition type paralleling -AgI behavior. The abrupt drop in conductivity in the LT region for P between 2.5–4.1 GPa of all compounds is at variance with the Arrhenius behavior observed for unperturbed ion migration implying the appearance of a second factor overriding the Arrhenius temperature dependence. Normal Arrhenius σT dependence prevails in both IT and HT regions with Qc values of 85–100 kJ mol−1 and 50–75 kJ mol−1, respectively. The higher conductivities at 0.4 GPa for TlBr and TlI relative to their 1 atm data and the increasing σ with P are in strong contrast to the normal σ-P behavior of TlCl. The dependence of activation volume ΔV on T for TlCl, i.e. ΔV>0, shows abnormally high values with a maximum at 500 K for P<3.0 GPa but reasonable ΔV values appear above 3.0 GPa. The ΔVT dependence for both TlBr and TlI with ΔV<0 is incompatible with an ion transport mechanism suggesting an electronic conduction process and implying an ionic–metallic transition at higher pressures. These contrasting conductivity features are discussed and interpreted in terms of electronegativity differences and bonding character rather than structure.  相似文献   

6.
The associating behaviour of the binary mixtures of ethylene glycol oligomer (EGO), i.e. ethylene glycol (EG), diethylene glycol (DEG) and poly(ethylene glycol)s (PEG200, PEG300, PEG400 and PEG600) with water (W), ethyl alcohol (EA) and 1,4-dioxane (DX) over the entire concentration range at 25 °C have been investigated through their accurately measured values of dielectric constant. The static dielectric constant εo, high frequency limiting dielectric constant ε, dielectric relaxation strength Δε, excess dielectric parameters εE0 and εE, effective Kirkwood correlation factor geff and corrective correlation factor gf of EGO–W, EGO–EA and EGO–DX mixtures were determined to obtain qualitative and quantitative information about the complex formation through H-bond in these systems. Most of the evaluated dielectric parameters of EG and DEG in different ‘cosolvents’ have different characteristics as compared to the PEG–cosolvent mixture. The observed linear and non-linear behaviour of Δε against EGO monomer unit mole fraction X confirms the variation in the homogeneous structures in their binary mixtures with concentration variation. Appearance of the maximum in εE0 against X plots indicates that a complex stable adduct is formed in the EGO–W mixtures at stoichiometric ratio 1:1.7 for lower oligomers but this ratio seems to be 1.7:1 for higher EGO molecules, which confirms that the EGO size and chain flexibility affects the complex formation between EGO and W. In case of EG–EA mixture 1:1 stoichiometric ratio form stable adduct whereas for higher EGO–EA, it is 3:1, at EGO monomer unit level. The complex formation behaviour of DEG–EA has entirely different characteristics when compared to the other studied EGO–EA mixtures. Although, 1,4-dioxane has weak polar behaviour dielectric properties of EG–DX and DEG–DX confirm the formation of stable adducts at the stoichiometric ratio 2:1 of EGO monomer unit mole fraction to the DX. For the higher EGO–DX mixtures, stable adduct forms at the stoichiometric ratio 9:1. Except DEG–EA mixtures, the EGO–W and EGO–EA form the complex with reduction in the effective number of dipoles. In EG–DX mixtures, the heterogeneous species form with a large reduction in the effective number of dipoles, which changes as the effective number of dipoles increases with the increase in monomer repeat units of EGO. Further the net electronic polarization in these binary mixtures increases due to heterogeneous interaction over the entire mixing concentration range.  相似文献   

7.
The dielectric relaxation time (τ) of binary mixtures of different molar concentrations of pyridine (C5H5N) and chlorobenzene (C6H5Cl) in benzene solution at different temperatures (25, 30, 35 and 40 °C) has been calculated by using standard microwave techniques and Gopala Krishna's single frequency (9.875 GHz) concentration variation method. The energy parameters (ΔHε, ΔFε, and ΔSε) for the dielectric relaxation process of the binary mixture containing 0.5 mol fraction of pyridine have been calculated at the respective temperatures. Comparisons have been made with the corresponding energy parameters for the viscous flow (ΔHη, ΔFη, and ΔSη). From the observations it is found that the dielectric relaxation process can be treated as the rate process. Based upon above studies, solute–solvent type of molecular associations arising from the interaction of chlorobenzene and benzene and pyridine and benzene molecules has been proposed. No solute–solute type of molecular association has been observed.  相似文献   

8.
The electrical conductances of the solutions of sodium bromide (NaBr) and sodium tetraphenylborate (NaBPh4) in 2-ethoxyethanol + water mixed solvent media containing 0.25, 0.50 and 0.75 mass fractions of 2-ethoxyethanol (w1) have been reported at 308.15, 313.15, 318.15, and 323.15 K. The conductance data have been analyzed by the 1978 Fuoss conductance–concentration equation in terms of the limiting molar conductance (Λ0), the association constant (KA) and the association diameter (R). Of the two electrolytes investigated here, sodium bromide exists essentially in the form of free ions in aqueous 2-ethoxyethanol solutions over the entire temperature range investigated. However, slight ionic association was observed for sodium tetraphenylborate—the extent of which increases with increasing amount of 2-ethoxyethanol in the present mixed solvent media. The solvations of the bromide ion and of the sodium ion were found to be gradually weakened as the 2-ethoxyethanol content of the medium increases. Furthermore, the limiting molar conductivity values of the two electrolytes increase as the temperature increases in all 2-ethoxyethanol + water mixtures which have been described by polynomial equations.  相似文献   

9.
It is known that the Korteweg–de Vries (KdV) equation is a geodesic flow of an L2 metric on the Bott–Virasoro group. This can also be interpreted as a flow on the space of projective connections on S1. The space of differential operators Δ(n)=∂n+u2n−2++un form the space of extended or generalized projective connections. If a projective connection is factorizable Δ(n)=(∂−((n+1)/2−1)p1)(∂+(n−1)/2pn) with respect to quasi primary fields pi’s, then these fields satisfy ∑i=1n((n+1)/2−i)pi=0. In this paper we discuss the factorization of projective connection in terms of affine connections. It is shown that the Burgers equation and derivative non-linear Schrödinger (DNLS) equation or the Kaup–Newell equation is the Euler–Arnold flow on the space of affine connections.  相似文献   

10.
《Physics letters. [Part B]》2002,550(3-4):147-153
Narrow structures in the range of a few MeV have been searched for in ppπ+ and ppπ invariant mass spectra (Mppπ+ and Mppπ) obtained from exclusive measurements of the ppppπ+π reaction at Tp=725,750 and 775 MeV using the PROMICE/WASA detector at CELSIUS. The selected reaction is particularly well suited for the search for dibaryon resonances decoupled from NN and/or NΔ. In the mass range 2020 MeV/c2<mdibaryon<2085 MeV/c2 no narrow structures could be identified on the 3σ level of statistical significance neither in Mppπ nor in Mppπ+ giving an upper limit (95% C.L.) for dibaryon production in this reaction of σ<20 nb.  相似文献   

11.
The current–voltage (IV) and capacitance–voltage (CV) characteristics of H-terminated Pb/p-Si/Al contacts fabricated by us have been measured in the temperature range of 77–300 K. The experimental values of the barrier height (BH) Φbo and the ideality factor n for the device range from 0.674 and 1.072 eV (at 300 K) to 0.352 and 2.452 eV (at 77 K), respectively. The ideality factors become larger with lowering temperature while the barrier height decreases. The Φbo(n) plot shows a linear dependence in the temperature range of 77–300 K that can be explained by the barrier inhomogeneity at the metal/semiconductor interface. The extrapolation of the linear Φbo(n) plot to n = 1 has given a homogeneous barrier height of approximately 0.713 eV for the Pb/p-Si(1 0 0) contact. A Φbo versus 1/T plot was drawn to obtain evidence of a Gaussian distribution of the BHs, and values of and σs = 80.5 mV for the mean BH and zero-bias standard deviation have been obtained from this plot, respectively. Then, a modified versus 1/T plot gives and A* as 0.828 eV and 54.89 A/cm2 K2, respectively. Furthermore, an average value of −0.687 meV/K for the temperature coefficient has been obtained, the value of −0.687 meV/K for hydrogen terminated p-type Si differs from those given for p-type Si without hydrogen termination in the literature.  相似文献   

12.
Characterization by Auger electron spectroscopy (AES) and Fourier transformation infrared spectroscopy (FTIR) confirms (Ta2O5)x(Al2O3)1−x alloys are homogeneous pseudo-binary alloys with increased thermal stability with respect to end member oxides, Ta2O5 and Al2O3. Capacitance–voltage (CV) and current density–voltage (JV) data as a function of temperate show that the Ta d-states of the alloys act as localized electron traps, and are at an energy approximately equal to the conduction band offset of Ta2O5 with respect to Si.  相似文献   

13.
The equations of state (EOS) and other thermodynamic properties of TiB2 are investigated using ab initio plane-wave pseudopotential density functional theory method. The obtained results are in good agreement with the experimental measured data and other theoretical calculated ones. Through the quasi-harmonic Debye model, in which the phononic effects are considered, the dependences of relative volume V/V0 on pressure P, cell volume V on temperature T, and Debye temperature Θ and specific heat CV on pressure P are successfully obtained.  相似文献   

14.
Young-Su Lee  Han-Ill Yoo   《Solid State Ionics》2002,150(3-4):373-382
Current (I)–voltage (V) characteristic under oxygen potential gradients was experimentally examined on single crystal BaTiO3−δ in its mixed ion/electron/hole regime at 1000 °C. The variation of I vs. V appears similar to that of an n/p junction, but with the limiting slope (dI/dV) approaching the maximum and minimum possible equilibrium conductances in the given oxygen potential gradient as increasing forward and reverse bias, respectively. This characteristic has been precisely traced theoretically by using the partial ionic and electronic conductivities of BaTiO3−δ as measured against uniform oxygen chemical potential in equilibrium state. The nonlinear characteristic is attributed to the redistribution of oxygen chemical potential that is caused by a non-vanishing gradient of the ionic transference number of the oxide under the given oxygen potential gradient. It is demonstrated that the bulk transport properties of a mixed conductor may be tailored by terminal voltage in a chemical potential gradient.  相似文献   

15.
Excitation spectra for dispersed VUV-fluorescence of NO (λfl=134–152 nm) were measured in the exciting-photon energy range between 16.9 and 24.8 eV using monochromatised synchrotron radiation at medium bandwidth of 29 meV. Fluorescence from A1Π (vi=0, 1, 2) vibronic NO+-levels and fluorescence from excited dissociation fragments NI (3s 2PJ) was observed simultaneously by recording the dispersed fluorescence with a monochromator–position-sensitive detector combination. The autoionisation of NO (c3Π) nℓλ (vR=0) Rydberg levels into the NO+ A1Π (vi=0, 1, 2) vibronic levels was observed vibrationally resolved. Different Beutler–Fano profiles for autoionisation of one Rydberg level NO (c3Π) nℓλ (vR=0) into the different vibronic NO+ A1Π (vi=0, 1, 2) levels are clearly visible. The dependence of the Beutler–Fano profiles on the quantum numbers n, , and λ of the Rydberg electron is discussed. For the direct photoionisation into the NO+ A1Π (vi=0, 1, 2) vibronic levels a non-Franck–Condon behaviour was observed.  相似文献   

16.
Excess molar volumes of {(1−xA)CH3(CH2)n−1OH+xACH3CN} for n=1, 2, 3 or 4 as a function of composition under atmospheric pressure at 288.15, 293.15, 298.15 and 303.15 K have been used to test the applicability of the Prigogine–Flory–Patterson theory. According to the model, interactional contribution is the most important one to explain the VmE behavior. Good agreement is only achieved for the mixtures containing methanol (C1) or 1-butanol (C4). For the mixtures containing ethanol (C2) and 1-propanol (C3), which show an S-shaped VmE behavior, the correlation fails.  相似文献   

17.
high resolution solid state NMR spectroscopy was employed to study structural properties of bis(diisopropoxyphosphorothioyl) diselenide 1 and bis(dineopentoxyphosphorothioyl) diselenide 2. The principal elements Tii of effective dipolar/chemical shift tensor were calculated from spinning sideband intensities employing the WIN-MAS program. The values of anisotropy and asymmetry parameters reflect the distortion of the selenium environment. It was found that the T33 component mostly contributes to changes in the isotropic chemical shifts. CP/MAS experiments were used to decide the assignment of space group by counting the number of crystallographically unique selenium centers in the unit cell. Crystals of diselenide 1 are triclinic, space group P with a=8.485(3) Å, b=8.508(1) Å, c=8.511(2) Å, =98.835(15)°, β=111.653(24)°, γ=93.524(21)°, V=559.5(3) Å3, Dc=1.544(2) g/cm3 and Z=1. Refinement using 2222 reflections for 157 variables gives R=0.037. Crystals of diselenide 2 are triclinic, space group P1 with a=9.1418(8) Å, b=9.1465(8) Å, c=9.9200(9) Å, =74.751(8)°, β=74.629(7)°, γ=82.216(7)°, V=769.7(1) Å3, Dc=1.365(2) g/cm3 and Z=1. Refinement using 3316 reflections for 297 variables gives R=0.0272.  相似文献   

18.
The present communication reports the experimental values of NMR spin-lattice relaxation time (T1) and dielectric relaxation time (τ) of piperidine, pyrrole, pyridine, diethylamine, triethylamine and pyrrolidine. The values of activation energy (ΔEA) obtained using dielectric relaxation time, have been correlated with calculated values of ΔEA obtained using Arrhenius equation of NMR relaxation time (T1) for pyridine, diethylamine and pyrrole. Authors have also established a correlation between the experimental values of NMR spin-relaxation time (T1) with its calculated values obtained using different equations of dielectric relaxation time (τ).  相似文献   

19.
Baryon Regge-pole contributions to u-channel helicity amplitudes for the γN→πN processes are derived, with attention to kinematic singularities and threshold conditions. An N, Nγ and Δδ Regge-exchange model is proposed that describes the backward scattering data on γp→πop and γp→π+n at high energy. The N and Nγ trajectories are found to be nearly degenerate, with residues in the ratio β(Nγ)/β(N)≈0.6. Structure in the differential cross sections is explained as dominance at small u giving way to dominance at large u. An isoscaar-isovector admixture for the γ-coupling to is required by the fits. The solution extrapolates through the mean γp→πop 180o differential cross section at intermediate energies, as required by duality.  相似文献   

20.
This work is a continuation of our studies of the liquid structure of polyhydric alcohols and their solutions. Dielectric parameters (relaxation time, average cluster dipole moment μ¯c, etc.) have been calculated for 1-hexanol in a wide temperature range in the framework of the cluster model proposed by Dissado and Hill; the values of apparent activation enthalpy ΔHDHexp have been evaluated. These data were compared with the parameters previously determined by us for 2,5-hexanediol and 1,2,6-hexanetriol; the ratio μ¯c/μ¯v was used for an estimation of the number of molecules in clusters. It was found that at all temperatures in the range 303–393 K, the μ¯c/μ¯v values for 1-hexanol are significantly larger than the values of average degree of association calculated from equilibrium dielectric data; the ΔHDHexp values for 1-hexanol are considerably lower than those for 2,5-hexanediol and 1,2,6-hexanetriol.  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