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 750 毫秒
1.
Apparent molar heat capacities Cp, φand apparent molar volumesVφ were determined for aqueous solutions of N, N - dimethylformamide andN , N - dimethylacetamide at temperatures from 278.15 to 393.15 K and at the pressure 0.35 MPa. The molalities investigated ranged from 0.015 mol ·kg  1to 1.0 mol · kg  1. We used a vibrating-tube densimeter (DMA 512P, Anton PAAR, Austria) to determine the densities and volumetric properties. Heat capacities were obtained using a twin fixed-cell, power-compensation, differential-output, temperature-scanning calorimeter (NanoDSC 6100, Calorimetry Sciences Corporation, Spanish Fork, UT, U.S.A.). The results were fit by regression to equations that describe the surfaces (Vφ,T , m) and (Cp, φ, T, m). Infinite dilution partial molar volumes V2oand heat capacitiesCp,2o were obtained over the range of temperatures by extrapolation of these surfaces to m =  0.  相似文献   

2.
Apparent molar heat capacities Cp, φand apparent molar volumesVφ were determined for aqueous solutions of 1-butanol, 2-butanol (both R andS isomers), isobutanol (2-methyl-1-propanol), and t -butanol (2-methyl-2-propanol) at temperatures from 278.15 K to 393.15 K and at the pressure 0.35 MPa. The molalities investigated ranged from 0.02 mol · kg  1to 0.5 mol · kg  1. We used a vibrating-tube densimeter (DMA 512P, Anton Paar, Austria) to determine the densities and volumetric properties. Heat capacities were obtained using a twin fixed-cell, power-compensation, differential-output, temperature-scanning calorimeter (NanoDSC 6100, Calorimetry Sciences Corporation, Provo, UT, U.S.A.). The results were fit by regression to equations that describe the surfaces (Vφ, T, m) and (Cp,φ, T, m). Infinite dilution partial molar volumesV2o and heat capacities Cp,2owere obtained over the range of temperatures by extrapolation of these surfaces to m =  0.  相似文献   

3.
We have measured the densities of aqueous solutions of l-methionine, l-methionine plus equimolal HCl, and l-methionine plus equimolal NaOH at temperatures 278.15  T/K  368.15, at molalities 0.0125  m/mol · kg−1  1.0 as solubilities allowed, and at p = 0.35 MPa using a vibrating tube densimeter. We have also measured the heat capacities of these solutions at 278.15  T/K  393.15 and at the same m and p using a twin fixed-cell differential temperature-scanning calorimeter. We used the densities to calculate apparent molar volumes Vϕ and the heat capacities to calculate apparent molar heat capacities Cp,ϕ for these solutions. We used our results and values from the literature for Vϕ(T, m) and Cp,ϕ(T, m) for HCl(aq), NaOH(aq), and NaCl(aq) and the molar heat capacity change ΔrCp,m(T, m) for ionization of water to calculate parameters for ΔrCp,m(T, m) for the two proton dissociations from protonated aqueous cationic l-methionine. We integrated these results in an iterative algorithm using Young’s Rule to account for the effects of speciation and chemical relaxation on Vϕ(T, m) and Cp,ϕ(T, m). This procedure yielded parameters for Vϕ(T, m) and Cp,ϕ(T, m) for methioninium chloride {H2Met+Cl(aq)} and for sodium methioninate {Na+Met(aq)} which successfully modeled our observed results. Values are given for ΔrCp,m, ΔrHm, pQa, ΔrSm, and ΔrVm for the first and second proton dissociations from protonated aqueous l-methionine as functions of T and m.  相似文献   

4.
Apparent molar heat capacities Cp, φand apparent molar volumesVφ were determined for aqueous solutions of α - and β -cyclodextrins at temperatures from 278.15 K to 393.15 K and at the pressure 0.35 MPa. The molalities investigated ranged from 0.008 mol · kg  1to 0.12 mol · kg  1forα -cyclodextrin and from 0.004 mol · kg  1to 0.014 mol · kg  1for β -cyclodextrin. We used a vibrating-tube densimeter (DMA 512P, Anton PAAR, Austria) to determine the densities and volumetric properties. Heat capacities were obtained using a twin fixed-cell, power-compensation, differential-output, temperature-scanning calorimeter (NanoDSC 6100, Calorimetry Sciences Corporation, Spanish Fork, UT, USA). Equations were fit by regression to our experimental (Vφ, T, m) and (Cp, φ,T , m) results. Infinite dilution partial molar volumes V2oand heat capacities Cp,2owere obtained over the range of temperatures by extrapolation of these surfaces to m =  0.  相似文献   

5.
In adiabatic low-pressure and dynamic calorimeters the temperature dependence of the standard molar heat capacity Cp, moof dibenzo- p -dioxin and 1,2,3,4-tetrachlorodibenzo- p -dioxin have been determined at temperatures in the range T =  5 K to T =  490 K: from T =  5 K to T =  340 K with an accuracy of about 0.2 per cent and with an accuracy of 0.5 per cent to 1.5 per cent between T =  340 K and T =  490 K. The temperatures, enthalpies, and entropies of melting of the above compounds have been determined. The experimental data were used to calculate the thermodynamic functions Cp, mo / R, Δ0THmo / (R·K), Δ0TSmo / R, and Φmo = Δ0TSmo  Δ0THmo / T(where R is the universal gas constant) in the range T   0 to T =  490 K. The isochoric heat capacity CV, mof both dioxins has been estimated over the range T   0 to Tfus. The effect of substitution of four hydrogen atoms by chlorine atoms on the lattice and atomic components of the isochoric heat capacity was considered.  相似文献   

6.
We determined apparent molar volumes V? at 278.15 ? (T/K) ? 368.15 and apparent molar heat capacities Cp,? at 278.15 ? (T/K) ? 393.15 at p = 0.35 MPa for aqueous solutions of tetrahydrofuran at m from (0.016 to 2.5) mol · kg?1, dimethyl sulfoxide at m from (0.02 to 3.0) mol · kg?1, 1,4-dioxane at m from (0.015 to 2.0) mol · kg?1, and 1,2-dimethoxyethane at m from (0.01 to 2.0) mol · kg?1. Values of V? were determined from densities measured with a vibrating-tube densimeter, and values of Cp,? were determined with a twin fixed-cell, differential, temperature-scanning calorimeter. Empirical functions of m and T for each compound were fitted to our V? and Cp,? results.  相似文献   

7.
We have measured the densities of aqueous solutions of isoleucine, threonine, and equimolal solutions of these two amino acids with HCl and with NaOH at temperatures 278.15  T/K  368.15, at molalities 0.01  m/mol · kg−1  1.0, and at the pressure 0.35 MPa using a vibrating tube densimeter. We have also measured the heat capacities of these solutions at 278.15  T/K  393.15 and at the same m and p using a twin fixed-cell differential temperature-scanning calorimeter. We used the densities to calculate apparent molar volumes Vϕ and the heat capacities to calculate apparent molar heat capacities Cp,ϕ for these solutions. We used our results and values from the literature for Vϕ(T, m) and Cp,ϕ(T, m) for HCl(aq), NaOH(aq), and NaCl(aq) and the molar heat capacity change ΔrCp,m(T, m) for ionization of water to calculate parameters for ΔrCp,m(T, m) for the two proton dissociations from each of the protonated aqueous cationic amino acids. We used Young’s Rule and integrated these results iteratively to account for the effects of equilibrium speciation and chemical relaxation on Vϕ(T, m) and Cp,ϕ(T, m). This procedure gave parameters for Vϕ(T, m) and Cp,ϕ(T, m) for threoninium and isoleucinium chloride and for sodium threoninate and isoleucinate which modeled our observed results within experimental uncertainties. We report values for ΔrCp,m, ΔrHm, pQa, ΔrSm, and ΔrVm for the first and second proton dissociations from protonated aqueous threonine and isoleucine as functions of T and m.  相似文献   

8.
We have measured the densities of aqueous solutions of serine, serine plus equimolal HCl, and serine plus equimolal NaOH at temperatures 278.15  T/K  368.15, molalities 0.01  m/mol · kg−1  1.0, and at the pressure p = 0.35 MPa, using a vibrating tube densimeter. We have also measured the heat capacities of these solutions at 278.15  T/K  393.15 and at the same m and p using a fixed-cell differential temperature-scanning calorimeter. We used the densities to calculate apparent molar volumes Vϕ and the heat capacities to calculate apparent molar heat capacities Cp,ϕ for these solutions. We used our results and values from the literature for Vϕ(T,m) and Cp,ϕ(T,m) for HCl(aq), NaOH(aq), and NaCl(aq) and the molar heat capacity change ΔrCp,m(T,m) for ionization of water to calculate ΔrCp,m(T,m) for proton dissociations from protonated aqueous cationic serine and from the zwitterionic form. We integrated these results in an iterative algorithm using Young’s rule to account for the effects of speciation and chemical relaxation on the observed Vϕ(T,m) and Cp,ϕ(T,m) of the solutions. This procedure yielded parameters for Vϕ(T,m) and Cp,ϕ(T,m) for serinium chloride {H2Ser+Cl(aq)} and for sodium serinate {Na+Gly(aq)} which successfully modeled our observed results. We have then calculated ΔrCp,m, ΔrHm, ΔrVm and pQa for the first and second proton dissociations from protonated aqueous serine as functions of T and m.  相似文献   

9.
We have measured the densities of aqueous solutions of glycine, glycine plus equimolal HCl, and glycine plus equimolal NaOH at temperatures 278.15  T/K  368.15, molalities 0.01  m/mol · kg−1  1.0, and at p = 0.35 MPa, using a vibrating tube densimeter. We have also measured the heat capacities of these solutions at 278.15  T/K  393.15 and at the same m and p using a fixed-cell differential scanning calorimeter. We used the densities to calculate apparent molar volumes Vϕ and the heat capacities to calculate apparent molar heat capacities Cp,ϕ for these solutions. We used our results and values of Vϕ(T, m) and Cp,ϕ(T, m) for HCl(aq), NaOH(aq), NaCl(aq) from the literature to calculate parameters for ΔrCp,m(T, m) for the first and second proton dissociations from protonated aqueous cationic glycine. We then integrated this value of ΔrCp,m(T, m) in an iterative algorithm, using Young’s Rule to account for the effects of speciation and chemical relaxation on the observed Vϕ and Cp,ϕ of the solutions. This procedure yielded parameters for Vϕ(T, m) and Cp,ϕ(T, m) for glycinium chloride {H2Gly+Cl(aq)} and sodium glycinate {Na+Gly(aq)} which successfully modeled our observed results. We have then calculated values of ΔrCp,m, ΔrHm, ΔrVm, and pQa for the first and second proton dissociations from protonated aqueous glycine as functions of T and m.  相似文献   

10.
We have measured the densities of aqueous solutions of alanine, alanine plus equimolal HCl, and alanine plus equimolal NaOH at temperatures 278.15  T/K  368.15, at molalities 0.0075  m/mol · kg−1  1.0, and at the pressure p = 0.35 MPa using a vibrating tube densimeter. We have also measured the heat capacities of these solutions at 278.15  T/K  393.15 and at the same m and p using a twin fixed-cell differential temperature-scanning calorimeter. We used the densities to calculate apparent molar volumes Vϕ and the heat capacities to calculate apparent molar heat capacities Cp,ϕ for these solutions. We used our results and values from the literature for Vϕ(T, m) and Cp,ϕ(T, m) for HCl(aq), NaOH(aq), and NaCl(aq) and the molar heat capacity change ΔrCp,m(T, m) for ionization of water to calculate parameters for ΔrCp,m(T, m) for the two proton dissociations from protonated aqueous cationic alanine. We integrated these results in an iterative algorithm using Young’s Rule to account for the effects of speciation and chemical relaxation on Vϕ(T, m) and Cp,ϕ(T, m). This procedure yielded parameters for Vϕ(T, m) and Cp,ϕ(T, m) for alaninium chloride {H2Ala+Cl(aq)} and for sodium alaninate {Na+Ala(aq)} which successfully modeled our observed results. Values are given for ΔrCp,m, ΔrHm, pQa, ΔrSm, and ΔrVm for the first and second proton dissociations from protonated aqueous alanine as functions of T and m.  相似文献   

11.
We determined apparent molar volumes V? at 298.15 ? (T/K) ? 368.15 and apparent molar heat capacities Cp,? at 298.15 ? (T/K) ? 393.15 for aqueous solutions of HIO3 at molalities m from (0.015 to 1.0) mol · kg?1, and of aqueous KIO3 at molalities m from (0.01 to 0.2) mol · kg?1 at p = 0.35 MPa. We also determined V? at the same p and at 298.15 ? (T/K) ? 368.15 for aqueous solutions of KI at m from (0.015 to 7.5) mol · kg?1. We determined Cp,? at the same p and at 298.15 ? (T/K) ? 393.15 for aqueous solutions of KI at m from (0.015 to 5.5) mol · kg?1, and for aqueous solutions of NaIO3 at m from (0.02 to 0.15) mol · kg?1. Values of V? were determined from densities measured with a vibrating-tube densimeter, and values of Cp,? were determined with a twin fixed-cell, differential temperature-scanning calorimeter. Empirical functions of m and T were fitted to our results for each compound. Values of Ka, ΔrHm, and ΔrCp,m for the proton ionization reaction of aqueous HIO3 are calculated and discussed.  相似文献   

12.
The temperature dependence of the standard molar heat capacity Cp, moof samples of crystalline tetraphenylphosphonium perchlorate and tetraphenylarsonium perchlorate was measured in an adiabatic low-pressure calorimeter between T =  4.8 K and T =  340 K and from T =  5.8 K to T =  340 K, respectively, mostly to within a precision of 0.2 per cent. For tetraphenylphosphonium perchlorate, an anomalous change of the heat capacity in the range T =  125 K to T =  185 K, probably arising from the excitation of hindered rotations of atomic groups, was found and its thermodynamic characteristics were determined. No such anomaly was observed for tetraphenylarsonium perchlorate. The data obtained were used to calculate the thermodynamic functions Cp, mo(T) / R, Δ0THmo / R·K, Δ0TSmo / R, and Φmo = Δ0TSmo  Δ0THmo / T(where R is the universal gas constant) of the compounds between T   0 and T =  340 K.  相似文献   

13.
The heat capacity Cp, mof NpO2was estimated for temperatures between 300 K and 1400 K. The Cp, mwas evaluated as a sum of three terms, phonon vibration Cph, m, dilation Cd, m, and Schottky specific heat Cs, m. The Cph, mand Cd,mwere calculated using the Debye temperature, Grüneisen constant and thermal expansion data obtained by high-temperature X-ray diffractometry. The coefficient of the linear thermal expansion (l.t.e.) for NpO2was given as a polynomial function up to T =  1573 K. The estimated Cp,mwas compared with that of previous studies. The present result at T =  300 K was 66.87 J · K  1· mol  1, which agreed well with previous results, 66.22 J · K  1· mol  1, measured by using calorimetry. The thermodynamic functions were given as a function of temperature.  相似文献   

14.
Apparent molar volumes Vϕ and apparent molar heat capacities Cp,ϕ were determined at the pressure 0.35 MPa for aqueous solutions of magnesium nitrate Mg(NO3)2 at molalities m = (0.02 to 1.0) mol · kg−1, strontium nitrate Sr(NO3)2 at m = (0.05 to 3.0) mol · kg−1, and manganese nitrate Mn(NO3)2 at m = (0.01 to 0.5) mol · kg−1. Our Vϕ values were calculated from solution densities obtained at T = (278.15 to 368.15) K using a vibrating-tube densimeter, and our Cp,ϕ values were calculated from solution heat capacities obtained at T = (278.15 to 393.15) K using a twin fixed-cell, differential, temperature-scanning calorimeter. Empirical functions of m and T were fitted to our results, and standard state partial molar volumes and heat capacities were obtained over the ranges of T investigated.  相似文献   

15.
The apparent molar volumes Vφ of glycine, alanine, valine, leucine, and lysine have been determined in aqueous solutions of 0.05, 0.5, 1.0 mol · kg−1 sodium dodecyl sulfate (SDS) and 1.0 mol · kg−1 cetyltrimethylammonium bromide (CTAB) by density measurements at T=298.15 K. The apparent molar volumes have also been determined for diglycine and triglycine in 1 mol · kg−1 SDS and CTAB solutions. These data have been used to calculate the infinite dilution apparent molar volumes V20 for the amino acids and peptides in aqueous SDS and CTAB and the standard partial molar volumes of transfer (ΔtrV2,m0) of the amino acids and peptides to these aqueous surfactant solutions. The linear correlation of V20 for a homologous series of amino acids has been utilized to calculate the contribution of the charged end groups (NH3+, COO), CH2 group and other alkyl chains of the amino acids to V20. The results on the partial molar volumes of transfer from water to aqueous SDS and CTAB have been interpreted in terms of ion–ion, ion–polar and hydrophobic–hydrophobic group interactions. The volume of transfer data suggests that ion–ion or ion–hydrophilic group interactions of the amino acids and peptides are stronger with SDS compared to those with CTAB. Comparison of the hydration numbers of amino acids calculated in the present studies with those in other solvents from literature shows that these numbers are almost the same at 1 mol · kg−1 level of the cosolvent/cosolute. Increasing molality of the cosolvent/cosolute beyond 1 mol · kg−1 lowers the hydration number of the amino acids due to increased interactions with the solvent and reduced electrostriction.  相似文献   

16.
The viscosities, η of mono-, di-, tri-saccharides and methylglycosides, viz., d(+)-xylose (XYL), d(?)-arabinose (ARA), d(?)-ribose (RIB), d(?)-fructose (FRU), d(+)-galactose (GAL), d(+)-mannose (MAN), d(+)-glucose (GLU), d(+)-melibiose (MEL), d(+)-cellobiose (CEL), d(+)-lactose monohydrate (LAC), d(+)-maltose monohydrate (MAL), d(+)-trehalose dihydrate (TRE), sucrose (SUC), d(+)-raffinose pentahydrate (RAF), α-methyl-d(+)-glucoside (α-Me-GLU), methyl-α-d-xylopyranoside (Me-α-XYL), and methyl-β-d-xylopyranoside (Me-β-XYL) in water and in (0.5, 1.0, 2.0, and 3.0) mol · kg?1 aqueous solutions of potassium chloride (KCl) have been determined at T = (288.15, 298.15, 308.15, and 318.15) K from efflux time measurements by using a capillary viscometer. Densities used to determine viscosities have been reported earlier. The viscosity data have been utilized to determine the viscosity B-coefficients employing the Jones–Dole equation at different temperatures. From these data, the viscosity B-coefficients of transfer, ΔtB have been estimated for the transfer of various saccharides/methylglycosides from water to aqueous potassium chloride solutions. The ΔtB values have been found to be positive, whose magnitude increases with the increase in concentration of potassium chloride in all cases. The dB/dT coefficients, pair, ηAB and triplet, ηABB viscometric interaction coefficients have also been determined. Gibbs free energies of activation and related thermodynamic parameters of activation of viscous flow have been determined employing Feakin’s transition-state theory. The signs and magnitudes of various parameters have been discussed in terms of solute–solute and solute–solvent interactions occurring in these solutions. The effect of substitution of –OH by methoxy group, –OCH3 has also been discussed.  相似文献   

17.
A heat-flow Calvet microcalorimeter was adapted for the measurement of sublimation enthalpies by the vacuum-drop method, with samples of masses in the range 1 mg to 5 mg. The electrically calibrated apparatus was tested by determining the enthalpies of sublimation of benzoic acid and ferrocene, at T =  298.15 K. The obtained results, ΔcrgHmo(C7H6O2)  =  (88.3  ±  0.5)kJ · mol  1and ΔcrgHmo(C10H10Fe) =  (73.3  ±  0.1)kJ · mol  1, are in excellent agreement with the corresponding values recommended in the literature. Subsequent application of the apparatus to the determination of the enthalpy of sublimation of η5-bis-pentamethylcyclopentadyenyl iron, at T =  298.15 K, led to ΔcrgHmo(C20H30Fe)  =  (96.8  ±  0.6)kJ · mol  1.  相似文献   

18.
A vibrating-tube densimeter (DMA 512P, Anton Paar, Austria) was used to investigate the densities and volumetric properties of aqueous potassium hydrogen phthalate (KHP) and potassium sodium phthalate (KNaP). Measurements were made at molalities m from (0.006 to 0.66)mol · kg  1, at temperatures from 278.15 K to 368.15 K and at the pressure 0.35 MPa. The densimeter was calibrated through measurements on pure water and on 1.0 mol · kg  1NaCl(aq). We also used a twin fixed-cell, power-compensation, differential-output, temperature-scanning calorimeter (NanoDSC 6100, Calorimetry Sciences Corporation, Spanish Fork, UT, U.S.A.) to measure solution heat capacities. This was accomplished by scanning temperature and comparing the heat capacities of the unknown solutions to the heat capacity of water. Apparent molar volumes Vφand apparent molar heat capacities Cp, φof the solutions were calculated and fit by regression to equations that describe the surfaces (Vφ, T, m) and (Cp, φ, T, m). Standard state partial molar volumesV2o and heat capacities Cp,2owere estimated by extrapolation to the m =  0 plane of the fitted surfaces. Previously determinedCp, φ for HCl(aq) and NaCl(aq) were used to obtain (ΔrCp, m, T, m) for the proton dissociation reaction of aqueous hydrogen phthalate. This (ΔrCp,m, T, m) surface was created by subtracting Cp,φfor KHP(aq) and for NaCl(aq) from the sum of Cp,φfor KNaP(aq) and for HCl(aq). Surfaces representing (ΔrHm, T, m) and (pQa, T, m), where pQadenotes the molality equilibrium quotient, were created by integration of our (ΔrCp,m, T, m) surface using values for (ΔrHm, m) and (pKa, m) at T =  308.15 K from the literature as integration constants.  相似文献   

19.
Density and ultrasound speed were measured accurately for diglycine + water, triglycine + water, diglycine + water-polyethylene glycol 400 (PEG400) and triglycine + water-PEG400 solutions at T = (293.15, 298.15, 303.15 and 308.15) K. The results were used in evaluating thermodynamic properties as apparent molar volumes (VØ) and apparent molar isentropic compressions (K) of diglycine and triglycine in water and in PEG400 solutions. Infinite dilution values of these parameters, VoØ, and Ko, were obtained from their plots as a function of molality by extrapolation and have been utilized in obtaining transfer volumes and transfer compressions at infinite dilution. All transfer volumes and transfer compressions were found to increase with increasing molality of PEG400. Apparent molar isobaric expansions were derived from the temperature dependence of VØ values at infinite dilution and at finite concentrations. All the results were interpreted in terms of solute (diglycine or triglycine) and co-solute (PEG400) and solvent (H2O) interactions.  相似文献   

20.
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   

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