scispace - formally typeset
Search or ask a question

Showing papers on "Standard molar entropy published in 2005"


Journal ArticleDOI
TL;DR: Desorption studies reveal that Pb2+ can be easily removed from carbon Nanotubes by altering the pH values of the solution using both HCl and HNO3, indicating that carbon nanotubes are a promising absorbent for wastewater treatment.

654 citations


Journal ArticleDOI
TL;DR: Adsorption mechanism studies revealed that the process was complex and followed both surface adsorption and particle diffusion and the rate-controlling parameter and effective diffusion coefficient were determined using the Reichenberg plot.

545 citations


Journal ArticleDOI
TL;DR: The Henry's constants of water, carbon dioxide, ethane, ethene, methane, oxygen, and nitrogen are computed in the ionic liquid 1-n-butyl-3-methylimidazolium hexafluorophosphate using test particle insertion and expanded ensemble Monte Carlo methods.
Abstract: The Henry's constants of water, carbon dioxide, ethane, ethene, methane, oxygen, and nitrogen are computed in the ionic liquid 1-n-butyl-3-methylimidazolium hexafluorophosphate ([bmim][PF6]) using test particle insertion and expanded ensemble Monte Carlo methods. The partial molar enthalpy and partial molar entropy of solvation are also computed for water, carbon dioxide, and oxygen. The results from the simulations are compared against experimental data from the literature. In addition, the accuracy and precision of the two methods in determining the Henry's constant are examined. Local organization of the ionic liquid around a solute molecule is analyzed, and the interactions responsible for the experimentally observed solubility trends are identified.

162 citations


Journal ArticleDOI
TL;DR: In this paper, the extended Henry's law was applied to correlate the solubility data, and the thermodynamic properties such as standard enthalpy, standard Gibbs free energy, standard entropy, and standard heat capacity of these two systems were obtained.
Abstract: The solubilities of CO2 in 1-butyl-3-methylimidazolium hexafluorophosphate and 1,1,3,3-tetramethylguanidium lactate at the temperatures ranging from (297 to 328) K and the pressures ranging from (0 to 11) MPa were determined. The extended Henry's law was applied to correlate the solubility data, and the thermodynamic properties such as standard enthalpy, standard Gibbs free energy, standard entropy, and standard heat capacity of these two systems were obtained. Experimental results showed that the solubilities of CO2 in TMGL are slightly higher than those in [bmim][PF6].

125 citations


Journal ArticleDOI
TL;DR: In this article, the solubility of CO2 in sulfonate ionic liquids (ILs), such as trihexyl (tetradecyl) phosphonium dodecylbenzenesulfonates ([P6, 6,6,14][C12H25PhSO3]) and trihexymethyl (Tetradecyl) phonium mesylate ([P 6, 6 6, 14][MeSO3]), was determined at temperatures ranging from (305 to 325) K and pressures ranging from 4 to 9 MPa.
Abstract: In this work, the solubility of CO2 in sulfonate ionic liquids (ILs), such as trihexyl (tetradecyl) phosphonium dodecylbenzenesulfonates ([P6,6,6,14][C12H25PhSO3]) and trihexyl (tetradecyl) phosphonium mesylate ([P6,6,6,14][MeSO3]), was determined at temperatures ranging from (305 to 325) K and pressures ranging from (4 to 9) MPa. It was found that the difference of the solubility of CO2 in two kinds of sulfonate ILs is not dramatic on the basis of molality. The solubility of CO2 in [P6,6,6,14][MeSO3] is higher than that in [P6,6,6,14][C12H25PhSO3]. The solubility data were correlated by means of the extended Henry's law, and the thermodynamic functions, such as the standard enthalpy, standard Gibbs free energy, and standard entropy, were obtained. The Henry's law constant for CO2 in all the investigated ionic liquids increases with increasing temperature.

94 citations


Journal ArticleDOI
TL;DR: In this paper, the partial molar enthalpy and entropy of CaMnO 3− δ were derived for 2.92 > 2.86, within the estimated uncertainty (twice the standard deviation of the mean).

87 citations


Journal ArticleDOI
TL;DR: In this paper, a large single crystal of barium tungstate (BaWO4) with dimensions of 22mmdiameter×80mm length was grown by the Czochralski method using an iridium crucible.
Abstract: A large single crystal of barium tungstate (BaWO4) with dimensions of 22-mmdiameter×80-mm length was grown by the Czochralski method using an iridium crucible. The melting point, molar enthalpy of fusion, and molar entropy of fusion of the crystal were determined to be 1775.10 K, 96913.80Jmol−1, and 54.60JK−1mol−1, respectively. The average linear thermal-expansion coefficients are αa=10.9526×10−6∕K, αb=10.8069×10−6∕K, and αc=35.1063×10−6∕K in the temperature range from 303.15 to 1423.15 K along the three respective crystallographic axes. The density of the crystal follows an almost linear decrease from 6.393×103 to 6.000×103kgm−3 when the temperature is increased from 303.15 to 1423.15 K. The measured specific heat are 115.56–130.96JK−1mol−1 in the temperature range of 323.15–1173.15 K. The thermal diffusion coefficient of the crystal was measured in the temperature range of 297.15–563.15 K. The calculated thermal conductivity is 2.256Wm−1K−1 along the [001] direction and 2.324Wm−1K−1 along the [100] dir...

55 citations


Journal ArticleDOI
TL;DR: In this paper, Lanthanum cobaltate and its strontium-substituted products have been prepared using a novel stearic acid sol-gel method at 600°C.

38 citations


Journal ArticleDOI
TL;DR: In this paper, the standard entropy of formation at 298 K, Δ f S o 2 9 8, was determined from measuring the heat capacity, C p, from near-absolute zero (2 K) to 300 K by the relaxation method.
Abstract: The thermodynamic properties of Mg 2 Zn 3 were investigated by calorimetry. The standard entropy of formation at 298 K, Δ f S o 2 9 8 , was determined from measuring the heat capacity, C p , from near-absolute zero (2 K) to 300 K by the relaxation method. The standard enthalpy of formation at 298 K, Δ f H o 2 9 8 , was determined by solution calorimetry in hydrochloric acid solution. The standard Gibbs energy of formation at 298 K, Δ f G o 2 9 8 , was determined from these data. The results obtained were as follows: Δ f H o 2 9 8 (Mg 2 Zn 3 ) = -69.80 ′ 20 kJ . mol - 1 ; Δ f S o 2 9 8 (Mg 2 Zn 3 ) = -10.95 ′ 1.80 J . K - 1 . mol - 1 ; Δ f G o 2 9 8 (Mg 2 Zn 3 ) = -66.55 ′ 20 kJ mol - 1 . The electronic contribution to the heat capacity of Mg 2 Zn 3 was found to be similar to pure magnesium, indicating that the density of states in the vicinity of the Fermi level follows the free-electron parabolic law.

34 citations


Journal ArticleDOI
De-Zhi Sun1, Shi-Bing Wang1, Mingzhi Song1, Xi-Lian Wei1, Bao-Lin Yin1 
TL;DR: In this paper, the interaction between α-CD and three alkyl trimethyl ammonium bromides, a homologues series of surfactants, in aqueous solutions have been investigated with titration microcalorimetry at 298.15 K.
Abstract: Interactions between α-CD and three alkyl trimethyl ammonium bromides, a homologues series of surfactants, in aqueous solutions have been investigated with titration microcalorimetry at 298.15 K. The results are discussed in the light of the amphiphilic interaction and the iceberg structure of water molecules existing around the hydrophobic tail of the surfactant. The stoichiometry of the host–guest complex changes from 1:1 to 2:1, as the number of carbon atoms (n) in the hydrophobic chain, CnH2n+1, increases from 8 to 14. All the complexes are quite stable, with the apparent experiential stability constants being β1 = 2.65 × 103 dm3-mol−1, β2 = 4.85 × 106 dm6-mol−2, β2 = 6.50 × 106 dm6-mol−2, respectively for n = 8, 12, 14. All the complexation processes are shown to be enthalpy driven, and the standard enthalpy effect (−ΔH0) increases while standard entropy change (ΔS0) decreases with elongation of the hydrophobic chain.

32 citations


Journal ArticleDOI
TL;DR: In this paper, a systematic study of the non-stoichiometry and high-temperature transport properties of the perovskite related mixed conductor Sr 3 Fe 2 O 6 + δ is presented.

Journal ArticleDOI
TL;DR: In this article, an ionic liquid (IL) was prepared by directly mixing InCl3 and 1-methyl-3-butylimidazolium chloride (BMIC) with molar ratio 1:1 under dry argon atmosphere.

Journal ArticleDOI
TL;DR: In this paper, the low-temperature heat capacity of uvarovite (Ca3Cr2Si3O12) was measured between 2 and 400 K, and thermochemical functions were derived from the results.
Abstract: The low-temperature heat capacity of uvarovite (Ca3Cr2Si3O12) was measured between 2 and 400 K, and thermochemical functions were derived from the results The measured heat-capacity curve shows a signiÞ cant lambda-shaped anomaly peaking at around 9 K The nature of this transition is unknown From our data, we suggest a standard entropy for uvarovite at 29815 K of 3209 ± 06 J/(mol·K)

Journal ArticleDOI
TL;DR: Inverse gas chromatography (IGC) has been used to measure the surface adsorption of some probes (nalkanes on a series (n-C5 to n-C8) cyclohexane and benzene) on NaY and CaY zeolites.

Journal ArticleDOI
TL;DR: In this article, the standard molar enthalpy of formation Δ f H m ∘ was determined as −(2585.2 −4.9) kJ −1 −1.3121(T/K)−2.

Journal ArticleDOI
TL;DR: In this paper, the carbonate complexation of curium(III) in aqueous solutions with high ionic strength was investigated below solubility limits in the 10-70 C temperature range using time-resolved laser-induced fluorescence spectroscopy (TRLFS).
Abstract: The carbonate complexation of curium(III) in aqueous solutions with high ionic strength was investigated below solubility limits in the 10-70 C temperature range using time-resolved laser-induced fluorescence spectroscopy (TRLFS). The equilibrium constant, K3, for the Cm(CO3)2- + CO32- Cm(CO3)33- reaction was determined (log K3 = 2.01 ± 0.05 at 25C, I = 3 M (NaClO4)) and compared to scattered previously published values. The log K3 value for Cm(III) was found to increase linearly with 1/T, reflecting a negligible temperature influence on the corresponding molar enthalpy change, rH3 = 12.2 ± 4.4 kJ mol-1, and molar entropy change, rS3 = 79 ± 16 J mol-1 K-1. These values were extrapolated to I = 0 with the SIT formula (rH3 = 9.4 ± 4.8 kJ mol-1, rS3 = 48 ± 23 J mol-1 K-1, log K3 = 0.88 ± 0.05 at 25C). Virtually the same values were obtained from the solubility data for the analogous Am(III) complexes, which were reinterpreted considering the transformation of the solubility-controlling solid. The reaction studied was found to be driven by the entropy. This was interpreted as a result of hydration changes. As expected, excess energy changes of the reaction showed that the ionic strength had a greater influence on rS3 than it did on rH3.

Journal ArticleDOI
TL;DR: The citrate-nitrate gel combustion route was used to prepare SrFe12O19(s) powder sample and the compound was characterized by X-ray diffraction analysis.

Journal ArticleDOI
TL;DR: In this paper, the heat capacities of 1,4-dichlorobenzene, 1, 4-dibromobenzenes, and 1, 3, 5-trichlorobenenzene were measured by adiabatic calorimetry.
Abstract: The heat capacities of 1,4-dichlorobenzene, 1,4-dibromobenzene, and 1,3,5-trichlorobenzene from 5 K to 380 K and 1,3,5-tribromobenzene from 5 K to 410 K were measured by adiabatic calorimetry. The experimental data were used to calculate the molar entropy and enthalpy values relative to 0 K. Apart from 1,4-dichlorobenzene, the substances do not show any solid−solid transitions. Molar enthalpies of fusion and melting-point temperatures were determined. The results, given in order, are (17 907 ± 15) J·mol-1 and (326.24 ± 0.03) K, (20 387 ± 15) J·mol-1 and (360.48 ± 0.03) K, (17 557 ± 35) J·mol-1 and (335.92 ± 0.03) K, and (21 721 ± 20) J·mol-1 and (394.96 ± 0.07) K.

Journal ArticleDOI
TL;DR: The low-temperature heat capacity of Magnesioferrite (MgFe2O4) was measured between 1.5 and 300 K, and thermochemical functions were derived from the results as mentioned in this paper.
Abstract: The low-temperature heat capacity of magnesioferrite (MgFe2O4) was measured between 1.5 K and 300 K, and thermochemical functions were derived from the results. No heat capacity anomaly was observed. From our data, we suggest a standard entropy (298.15 K) for magnesioferrite of 120.8±0.6 J mol−1 K−1, which is about 2.4 J mol−1 K−1 higher than previously reported calorimetric studies; but is in rough agreement with predictions from sets of internally consistent thermodynamic data.

Journal ArticleDOI
TL;DR: In this paper, aqueous solutions of alanine and phenylalanine were prepared for five different molar concentrations of the respective amino acid and only one relaxation peak was observed in this frequency range.
Abstract: Dielectric relaxation measurements on aqueous solutions of alanine and phenylalanine were carried out using time domain reflectometry (TDR) at 25, 30, 35, and 40 °C in the frequency range from 10 MHz to 20 GHz. Aqueous solutions of alanine and phenylalanine are prepared for five different molar concentrations of the respective amino acid. For all the solutions considered, only one relaxation peak was observed in this frequency range. The relaxation peaks shift to lower frequency with an increase in alanine and phenylalanine molar concentration. The molar enthalpy of activation and molar entropy of activation show endothermic interactions.

Journal ArticleDOI
TL;DR: In this article, the standard entropy S 0 (298.15 K) of PuPO 4 was derived and a semi-empirical method was used to describe the total entropy as the sum of the lattice entropy S lat and the excess entropy S exs.

Journal ArticleDOI
TL;DR: In this paper, the formation thermodynamic properties of crystalline and aqueous inosine have been determined by using a combination of calorimetric techniques, and the standard molar enthalpy of formation Δ f H m ∘ is −(847.9 ± 4.5 ) kJ · mol - 1.

Journal ArticleDOI
TL;DR: In this paper, the standard Gibbs energies of reactions proceeding in the C-O system are calculated per mole of carbon dioxide, per mule of the reaction product, and per moles of atoms of the initial substances.
Abstract: The standard Gibbs energies of reactions proceeding in the C-O system are calculated per mole of carbon dioxide, per mole of the reaction product, and per mole of atoms of the initial substances. A comparative analysis of the calculation results is made.

Journal ArticleDOI
TL;DR: In this article, the standard entropy of formation at 298 k, Δ f S 298 o, of SnMg2 has been determined from measuring the heat capacities (Cp) from near absolute zero (2 k) to 300 k by the relaxation method.

Journal ArticleDOI
TL;DR: In this article, the equilibrium constants for 2-hexane, pentanone, and hexanone reactions were measured at several temperatures and the results showed that these values decrease monotonically with increasing value of the number of carbons and trend towards a limiting value of ≈030 for NC ≥ 8.

Journal Article
TL;DR: In this article, the molar entropy and enthalpy of the same one-component system are visualized as functions of temperature by numerical results of a Debye model, and a proper extrapolation does not cause an entropy catastrophe as claimed in Kauzmann's paradox, since the entropy difference between the undercooled melt and the corresponding crystals must be taken into account, and the respective entropies in both states are not connected by an isothermal process.
Abstract: The molar entropy, S, and enthalpy (energy), H, of crystals, glasses and melts of the same one-component systems have been suitably visualized including the transformation from the melt into a glass or crystallization. For the temperature T → 0 K the enthalpy and entropy of the glass are larger by ΔH 0 and ΔS 0 as compared to the stable crystal. The S and H functions of glasses correspond to a simple continuation of these functions from the molten state to lower temperatures. Crystallization occurs as a spontaneous process under production of entropy. Extrapolating the entropy of the molten and crystalline states from the melting range to lower temperatures, which is the basis of Kauzmann's paradox, is ambiguous and misleading, as the extrapolated data deviate considerably from the experimental temperature dependencies of S of glasses and crystals. A proper extrapolation does not cause an entropy catastrophe as claimed in Kauzmann's paradox, since the enthalpy difference between the undercooled melt and the corresponding crystals must be taken into account, and the respective entropies in both states are not connected by an isothermal process. The molar entropy and enthalpy are visualized as functions of temperature by numerical results of a Debye model. The molar entropy is a universal function of the ratio T/T D , wherein T D is the Debye temperature of the well known specific heat capacity, C C . Between 0 K and T D the entropy increases by 1.36 x 3R 4R irrespective of T D . Above T D it increases approximately as 3R × ln(T/T D ). The entropy capacity, C D /T, scales with 1/T D and the enthalpy with T D , both considered as functions of T/T D . The entropy capacity shows a maximum of 2.033 x 3R/T D for T/T D = 0.28.

Journal ArticleDOI
TL;DR: In this article, the melting point, molar enthalpy, and molar entropy of 2,4-dinitrobenzaldehyde were determined to be (344.91 +/- 0.15) K, (21,177 +/- 14) J (.) mol(-1), and (61.40 ± 0.07) J(.) K-1 (.)

Journal ArticleDOI
TL;DR: In this paper, the heat capacity of monoclinic gadolinium sesquioxide was measured by adiabatic calorimetry in the temperature range (4 to 380) K.

Journal ArticleDOI
TL;DR: The low temperature heat capacities of N-(2-cyanoethyl)aniline were measured with an automated adiabatic calorimeter over the temperature range from 83 to 353 K as discussed by the authors.

Journal Article
JZ Yang, QG Zhang, M Huang, F Xue, SL Zang 
TL;DR: In this article, a new theoretical model of ionic liquid, that is an interstice model, was applied to calculating the thermal expansion coefficient of BMIInCl4, and the results showed that there is much reasonableness for the inter-stice model of the ionic liquids.