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 62 毫秒
1.
2.
The electromotive force (e.m.f.)E of the cellPt|H2(p)|HCl(m) in Z|AgCl|Ag in {glycerol (G)  +  water (W)} solvents, Z  =  (G  +  W), up to glycerol mass fraction wG =  0.7 has been measured within the temperature range from 273.15 K to 313.15 K at HCl molalities up to 0.1mol · kg  1. On this basis, the standard molar e.m.f. Em  values pertaining to such solvent mixtures have been obtained, and have been combined with sparse literature data for optimization. At glycerol mass fractions up towG   0.5, at constant temperature,Em   shows a linear dependence on the glycerol mole fractionxG and, in parallel, Ec  (on the amount-of-substance concentration scale) shows linear dependence on the glycerol volume fractionϕG . The primary medium effect upon HCl, defined as the difference (Ec  )W   (Ec  )Z, has been considered as a function of the water volume fraction ϕWin terms of Feakins and French’s theory: this would lead to a primary hydration number n(hydr)  =  2.4 for HCl, in good agreement with previous results obtained with solvents other than (glycerol  +  water). In this connection, some basic methodological aspects are discussed. Ancillary values of the densities ρZof the relevant solvent mixtures, which were hitherto unavailable and are necessary for the data processing leading toEm   , have also been measured.  相似文献   

3.
New bimetallic complex salts corresponding to the formulation [Ni(L)][MCl4] have been synthesized by the facile reaction between [Ni(L)](ClO4)2 and [MCl2(PPh3)2] in high yields {where M = Co(II), Zn(II), Hg(II) and L = 3,7-bis(2-aminoethyl)-1,3,5,7-tetraazabicyclo(3.3.1)nonane}. The complexes were characterized by IR, electronic spectra, TGA/DSC, magnetic moment and conductivity measurements. The X-ray crystal structure for [Ni(L)][CoCl4] clearly establishes the cationic–anionic interaction. It crystallizes in the space group P1 with unit cell dimensions a = 7.1740(15) Å, b = 8.1583(16) Å and c = 8.3102(16) Å. A square-planar geometry is evident for the [Ni(L)]2+ cation while the anion is found to be tetrahedral. A two-step thermolytic pattern is observed in the pyrolysis of the bimetallic complex salts.  相似文献   

4.
Precise vapor pressure data for pure acetonitrile and (LiBr + acetonitrile) are given for temperatures ranging from T=(298.15 to 343.15) K. The molality range is from m=(0.0579 to 0.8298) mol · kg−1. The osmotic coefficients are calculated by taking into account the second virial coefficient of acetonitrile. The parameters of the extended Pitzer ion interaction model of Archer and the mole fraction-based thermodynamic model of Clegg–Pitzer are evaluated. These models accurately reproduce the available osmotic coefficients. The parameters of the extended Pitzer ion interaction model of Archer are used to calculate the mean molal activity coefficients.  相似文献   

5.
The apparent molar heat capacities Cp, φ  and apparent molar volumes Vφ  of Y2(SO4)3(aq), La2(SO4)3(aq), Pr2(SO4)3(aq), Nd2(SO4)3(aq), Eu2(SO4)3(aq), Dy2(SO4)3(aq), Ho2(SO4)3(aq), and Lu2(SO4)3(aq) were measured at T =  298.15 K and p =  0.1 MPa with a Sodev (Picker) flow microcalorimeter and a Sodev vibrating-tube densimeter, respectively. These measurements extend from lower molalities of m =  (0.005 to 0.018) mol ·kg  1to m =  (0.025 to 0.434) mol ·kg  1, where the upper molality limits are slightly below those of the saturated solutions. There are no previously published apparent molar heat capacities for these systems, and only limited apparent molar volume information. Considerable amounts of the R SO4 + (aq) and R(SO4)2  (aq) complexes are present, where R denotes a rare-earth, which complicates the interpretation of these thermodynamic quantities. Values of the ionic molar heat capacities and ionic molar volumes of these complexes at infinite dilution are derived from the experimental information, but the calculations are necessarily quite approximate because of the need to estimate ionic activity coefficients and other thermodynamic quantities. Nevertheless, the derived standard ionic molar properties for the various R SO4 + (aq) and R(SO4)2  (aq) complexes are probably realistic approximations to the actual values. Comparisons indicate that Vφ  {RSO4 + , aq, 298.15K}  =   (6  ±  4)cm3· mol  1and Vφ  {R(SO4)2  , aq, 298.15K}  =  (35  ±  3)cm3· mol  1, with no significant variation with rare-earth. In contrast, values of Cp, φ  { RSO4 + , aq, 298.15K } generally increase with the atomic number of the rare-earth, whereas Cp, φ  { R(SO4)2  , aq, 298.15K } shows a less regular trend, although its values are always positive and tend to be larger for the heavier than for the light rare earths.  相似文献   

6.
The samples of dibarium magnesium orthoborate Ba2Mg(BO3)2 were synthesized by solid-state reaction. The X-ray diffraction (XRD) patterns and Raman spectra of the samples were collected. Electronic structure and vibrational spectroscopy of Ba2Mg(BO3)2 were systematically investigated by first principle calculation. A direct band gap of 4.4 eV was obtained from the calculated electronic structure results. The top valence band is constructed from O 2p states and the low conduction band mainly consists of Ba 5d states. Raman spectra for Ba2Mg(BO3)2 polycrystalline were obtained at ambient temperature. The factor group analysis results show the total lattice modes are 5Eu + 4A2u + 5Eg + 4A1g + 1A2g + 1A1u, of which 5Eg + 4A1g are Raman-active. Furthermore, we obtained the Raman active vibrational modes as well as their eigenfrequencies using first-principle calculation. With the assistance of the first-principle calculation and factor group analysis results, Raman bands of Ba2Mg(BO3)2 were assigned as Eg (42 cm−1), A1g (85 cm−1), Eg (156 cm−1), Eg (237 cm−1), A1g (286 cm−1), Eg (564 cm−1), A1g (761 cm−1), A1g (909 cm−1), Eg (1165 cm−1). The strongest band at 928 cm−1 in the experimental spectrum is assigned to totally symmetric stretching mode of the BO3 units.  相似文献   

7.
The potential differences E of the cells Pt|H2|H2Ph(m1)  +  KHPh(m2)  +  KCl(m3) in Z|AgCl|Ag and Pt|H2|H2Ph(m1)  +  KHPh(m2)  +  KCl(m3) in Z|Hg2Cl2|Hg have been measured at T =  298.15 K in mixtures Z =  (W + S) of water (W) with cosolvents S =  propylene carbonate (PC) and S =  ethylene carbonate (EC), to determine the first ionization constants K of the o -phthalic acid H2Ph(benzene-1,2-dicarboxylic acid), which are indispensable for the determination of primary pH-metric standards based on the potassium hydrogen phthalate buffer (KHPh) in such solvent mixtures. The value of K is seen to decrease progressively with increasing mass fraction wsof the organic cosolvent, as with all of the other cosolvents studied earlier, but no simple relationship with the cosolvent permittivity is discernible. Since the required values of the standard potential difference Eoof the second cell were hitherto missing, they have now been obtained based on potential difference measurements of the cell Pt|H2|HCl(m) in Z|Hg2Cl2|Hg. The primary medium effect (EWo  EZo, by Owen’s definition) upon HCl in water-rich mixtures Z is seen to increase linearly with increasing ws, as in earlier investigations. In this comparative context, the slope of the primary medium effect against wsplots for the aprotic cosolvents increases regularly with decreasing permittivity, whereas for the protic (alcoholic) cosolvents the slope is ill-defined.  相似文献   

8.
Excess molar volumes VmEof {di- n -butyl ether (DBE)  +  a monofunctional organic compound} have been determined atT =  298.15 K over the whole composition range by means of a vibrating-tube densimeter. TheVmE values were either positive (propylamine, or butylamine, or acetone, or tetrahydrofuran  +  DBE) or negative (methanol, or butanol, or diethyl ether, or cyclopentanone, or acetonitrile  +  DBE). Markedly asymmetric VmEcurves were displayed by (DBE  +  methanol) and (DBE  +  acetonitrile). Partial molar volumes __ Vmoat infinite dilution in DBE, both from this work and the literature, were analysed in terms of an additivity scheme, and the group contributions thus obtained were discussed and compared with analogous results in water. DBE revealed a greater capability of distinguishing between polar and non-polar solutes, as well as in discriminating differently shaped molecules (unbranched, branched, cyclic). The limiting slopes of apparent excess molar volumes are evaluated and briefly discussed in terms of solute–solute and solute–solvent interactions.  相似文献   

9.
A new amino acid ionic liquid (AAIL) [C3mim][Val] (1-propyl-3-methylimidazolium valine) was prepared by the neutralization method. Using the solution-reaction isoperibol calorimeter, molar solution enthalpies of the ionic liquid [C3mim][Val] with known amounts of water and with different concentrations in molality were measured at T = 298.15 K. In terms of standard addition method (SAM) and Archer’s method, the standard molar enthalpy of solution for [C3mim][Val] without water, ΔsHm = (−55.7 ± 0.4) kJ · mol−1, was obtained. The hydration enthalpy of the cation [C3mim]+, ΔH+ ([C3mim]+) = −226 kJ · mol−1, was estimated in terms of Glasser’s theory. Using the RD496-III heat conduction microcalorimeter, the molar enthalpies of dilution, ΔDHm(mi  mf), of aqueous [C3mim][Val] with various values of molality were measured. The values of ΔDHm(mi  mf) were fitted to Pitzer’s ion-interaction model and the values of apparent relative molar enthalpy, φL, calculated using Pitzer’s ion-interaction model.  相似文献   

10.
Molar calorimetric enthalpy changes ΔrHm(cal) have been measured for the biochemical reactions {cAMP(aq) + H2O(l)=AMP(aq)} and {PEP(aq) + H2O(l)=pyruvate(aq) + phosphate(aq)}. The reactions were catalyzed, respectively, by phosphodiesterase 3,5-cyclic nucleotide and by alkaline phosphatase. The results were analyzed by using a chemical equilibrium model to obtain values of standard molar enthalpies of reaction ΔrHm for the respective reference reactions {cAMP(aq) + H2O(l)=HAMP(aq)} and {PEP3−(aq) + H2O(l)=pyruvate(aq) + HPO2−4(aq)}. Literature values of the apparent equilibrium constants K for the reactions {ATP(aq)=cAMP(aq) + pyrophosphate(aq)}, {ATP(aq) + pyruvate(aq)=ADP(aq) + PEP(aq)}, and {ATP(aq) + pyruvate(aq) + phosphate(aq)=AMP(aq) + PEP(aq) + pyrophosphate(aq)} were also analyzed by using the chemical equilibrium model. These calculations yielded values of the equilibrium constants K and standard molar Gibbs free energy changes ΔrGm for ionic reference reactions that correspond to the overall biochemical reactions. Combination of the standard molar reaction property values (K, ΔrHm, and ΔrGm) with the standard molar formation properties of the AMP, ADP, ATP, pyrophosphate, and pyruvate species led to values of the standard molar enthalpy ΔfHm and Gibbs free energy of formation ΔfGm and the standard partial molar entropy Sm of the cAMP and PEP species. The thermochemical network appears to be reasonably well reinforced and thus lends some confidence to the accuracy of the calculated property values of the variety of species involved in the several reactions considered herein.  相似文献   

11.
A flow mixing calorimeter followed by a vibrating-tube densimeter has been used to measure excess molar enthalpies HmE and excess molar volumes VmE of {xC4H10+(1−x)SF6}. Measurements over a range of mole fractions x have been made in the supercritical region at the pressure p=6.00 MPa and at seven temperatures in the range T=311.25 K to T=425.55 K. The HmE(x) measurements at T=351.35 K were found to exhibit an unusual double maximum. Measurements at all temperatures are compared with the Patel–Teja equation of state with the parameters determined by solving a cubic equation as recommended, and also with parameters determined by the method suggested by Valderamma and Cisternas who proposed equations which are a function of the critical compression factor. The overall fit to the HmE and VmE measurements obtained using Valderamma and Cisternas equations was found to be better than that obtained using the parameters according to the method suggested by Patel and Teja.  相似文献   

12.
Densities of binary mixtures of N,N-dimethylacetamide (DMA) with water (H2O) or water-d2 (D2O) were measured at the temperatures from T=277.13 K to T=318.15 K by means of a vibrating-tube densimeter. The excess molar volumes VmE, calculated from the density data, are negative for the (H2O + DMA) and (D2O + DMA) mixtures over the entire range of composition and temperature. The VmE curves exhibit a minimum at x(DMA)≅0.4. At each temperature, this minimum is slightly deeper for the (D2O + DMA) mixtures than for the corresponding (H2O + DMA) mixtures. The difference between D2O and H2O systems becomes smaller when the temperature increases. The VmE results were correlated using a modified Redlich–Kister expansion. The partial molar volume of DMA plotted against x(DMA) goes through a sharp minimum in the water-rich region around x(DMA)≅0.08. This minimum is more pronounced the lower the temperature and is deeper in D2O than in H2O at each temperature. Again, the difference becomes smaller as the temperature increases. The excess expansion factor αE plotted against x(DMA) exhibit a maximum in the water rich region of the mole fraction scale. At each temperature, this maximum is higher for the (D2O + DMA) mixtures than for the corresponding (H2O + DMA) mixtures, and the difference becomes smaller as the temperature increases. At its maximum, αE can be even more than 25 per cent of total value of the cubic expansion coefficient α in the (H2O + DMA) and (D2O + DMA) mixtures.  相似文献   

13.
The excess molar volumes VmE at T=298.15 have been determined in the whole composition domain for (2-methoxyethanol + tetrahydrofuran + cyclohexane) and for the parent binary mixtures. Data on VmE are also reported for (2-ethoxyethanol + cyclohexane). All binaries showed positive VmE values, small for (methoxyethanol + tetrahydrofuran) and large for the other ones. The ternary VmE surface is always positive and exhibits a smooth trend with a maximum corresponding to the binary (2-methoxyethanol + cyclohexane). The capabilities of various models of either predicting or reproducing the ternary data have been compared. The behaviour of VmE and of the excess apparent molar volume of the components is discussed in both binary and ternary mixtures. The results suggest that hydrogen bonding decreases with alcohol dilution and increases with the tetrahydrofuran content in the ternary solutions.  相似文献   

14.
A flow mixing calorimeter and a vibrating-tube densimeter have been used to measure excess molar enthalpies HmE and excess molar volumes VmE of {xC2H6 +  (1   x)SF6 }. Measurements over a range of mole fractions x have been made at T =  305.65 K and T =  312.15 K and at the pressures (3.76, 4.32, 4.88 and 6.0) MPa. The pressure 3.76 MPa is close to the critical pressure of SF6, the pressure 4.88 MPa is close to the critical pressure of C2H6, and the pressure 4.32 MPa is midway between these values. The measurements are compared with the Patel–Teja equation of state which reproduces the main features of the excess function curves as well as it does for similar measurements on {xCO2 +  (1   x)C2H6 }, {xCO2 +  (1   x)C2H4 } and {xCO2 +  (1   x)SF6 }.  相似文献   

15.
A flow mixing calorimeter followed by a vibrating-tube densimeter has been used to measure excess molar enthalpies HmE and excess molar volumesVmE of {xC3H8 +  (1   x)SF6}. Measurements over a range of mole fractionsx have been made at the pressure p =  4.30 MPa at eight temperatures in the rangeT =  314.56 K to 373.91 K, in the liquid region at p =  3.75 MPa andT =  314.56 K, in the two phase region at p =  3.91 MPa andT =  328.18 K, and in the supercritical region at p =  5.0 MPa andT =  373.95 K. The measurements are compared with results from the Patel–Teja equation of state which reproduces the main features of the excess function curves as well as it does for similar measurements on{xCO2 +  (1   x)C2H6} ,{xCO2 +  (1   x)C2H4} and{xCO2 +  (1   x)SF6} reported previously.  相似文献   

16.
Molar enthalpies of dilution ΔdilHmofNa2CO3(aq) were measured from molality m =  1.45 mol · kg  1to m =  0.008 mol · kg  1at seven temperatures from T =  298 K toT =  523 K at the pressure p =  7 MPa, and at four temperatures fromT =  371 K to T =  523 K at the pressurep =  40 MPa. Molar enthalpies of dilutionΔdilHm of NaHCO3(aq) were measured fromm =  0.98 mol · kg  1tom =  0.007 mol · kg  1at the same temperatures and pressures. Hydrolysis and ionization equilibria contribute substantially to the measured enthalpies under many of the conditions of this study. Explicit consideration of these reactions, using thermodynamic quantities from previous studies, facilitates a quantitative representation of apparent molar enthalpies, activity coefficients, and osmotic coefficients with the Pitzer ion-interaction treatment over the ranges of temperature, pressure, and molality of the experiments.  相似文献   

17.
A nonmagnetic compound, [NO2BzPy][Cu(mnt)2] (mnt2? = maleonitriledithiolate; NO2BzPy+ = 1-(4′-nitrobenzyl)pyridinium), is isostructural with [NO2BzPy][Ni(mnt)2], which is a quasi-1D spin system and exhibits a spin-Peierls-like transition with J = 192 K in the gapless state and spin energy gap = 738 K in the dimerization state, respectively. Further, five nonmagnetic impurity doped compounds [NO2BzPy][CuxNi1?x(mnt)2] (x = 0.04–0.74) were prepared, and their crystal structures as well as magnetic properties were investigated. The nonmagnetic doping causes the suppression of the spin transition with an average rate of 139(13) K/percentage of dopant concentration, and the transition collapse is estimated at around x > 0.5.  相似文献   

18.
Isothermal (vapour  +  liquid) equilibria were measured for (trichloromethane  +  tetrahydropyran or piperidine) at T =  333.15 K and {1-bromo-1-chloro-2,2,2-trifluoroethane (halothane)  +  tetrahydropyran or piperidine} atT =  323.15 K with a circulation still. The results were verified by effective statistical procedures and used to calculate activity coefficients and excess molar Gibbs free energiesGmE . Excess molar enthalpiesHmE for these mixtures were determined at T =  298.15 K by means of an isothermal CSC microcalorimeter equipped with recently reconstructed flow mixing cells. Reliable performance of the calorimetric setup was proved by the good agreement of HmEfor (hexane  +  cyclohexane), (2-propanone  +  water), and (methanol  +  water), with the best literature results. The trichloromethane- or halothane-containing mixtures exhibit strong negative deviations from Raoult’s law and are highly exothermic, thus indicating that complex formation via hydrogen bonding is a governing nonideality effect. A close similarity in the behaviour of corresponding mixtures with trichloromethane and halothane is observed, but for halothane-containing mixtures,GmE and HmEare consistently more negative, confirming that halothane is a more powerful proton donor than chloroform.  相似文献   

19.
Excess molar volumes VmEof (1,2-propanediol  +  water) and (1,2-butanediol  +  water) were measured at temperatures of (288.15, 298.15, and 308.15) K and at pressures of (0.1, 20, 40, and 60) MPa with a densimeter, model DMA 512p from Anton Paar. Values of VmEwere negative for all the mixtures studied over the whole concentration range and for all temperatures and pressures. Results were correlated by polynomial equations of Redlich and Kister and of Myers and Scott.  相似文献   

20.
Excess molar enthalpies HmEatT =  298.15 K are reported for (N -methyl-2-pyrrolidinone  +  chlorobenzene, or 1,2-dichlorobenzene, or 1,3-dichlorobenzene, or 1,2,4,-trichlorobenzene). The values ofHmE were obtained by using the flow calorimetric method. All the mixtures, over the whole composition range, are formed exothermically. The HmEresults are discussed in terms of the NRTL and UNIQUAC models.  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