Elsevier

Thermochimica Acta

Volume 561, 10 June 2013, Pages 1-13
Thermochimica Acta

Thermodynamic study on some alkanediol solutions: Measurement and modeling

https://doi.org/10.1016/j.tca.2013.03.010Get rights and content

Highlights

Abstract

The densities ρ and viscosities η of 1,2-ethanediol with 1,2-propanediol or 1,3-propanediol, and 1,2-propanediol with 1,3-propanediol binary liquid mixtures over the entire concentration range at temperatures (298.15 to 308.15) K with 5 K interval were measured. The experimental data were used to calculate the excess molar volume VmE, partial molar volume V¯m,i, partial molar volume at infinite dilution V¯i, apparent molar volume Vφi, coefficient of thermal expansion αp, excess coefficient of thermal expansion αpE, excess viscosity ηE, excess Gibbs energy of activation ΔG*E, and other thermodynamic parameters. A Redlich–Kister equation and Prigogine–Flory–Patterson (PFP) model was applied to correlate the excess molar volume results. Moreover, the viscosity data were correlated with the Grunberg–Nissan, Tamura–Kurata, Hind–Ubbelohde and Katti–Chaudhary equations. Good agreement was found between experimental data and modeling results.

Introduction

Most of liquid mixtures are non ideal and indicates particular behavior. In the last decade, large number of researches has been done on interpretation of non ideality behavior of solutions. In comparison to gas and solid, liquids are much more complicated state of the matter in nature because a diverse kind of interactions can be influenced in solution behavior [1]. The thermophysical properties of solutions including density, viscosity and their corresponding volumetric and viscosimetric properties such as excess molar volume and excess viscosity are useful in the study of molecular dynamic and molecular interaction [2], [3], [4]. These solution properties not only depended on solute-solute, solvent-solvent, and solute-solvent interactions, but also on the structural effect arising from interstitial accommodation owing to difference in volume and free volume between components present in the solution [5]. Moreover, investigation on the thermophysical properties is essential for designing, testing, and extending of theoretical models for fluid mixtures [6], [7], [8], [9]. From the industrial view point, many processes require accurate thermodynamic and transport values of liquids which are indispensable for many practical problem concerning heat transport, mass transport, fluid flow, and pipeline systems [10], [11].

Alkanediols are the class of materials with the ability to form intra molecular hydrogen bonding within their molecules and also the tendency to form hydrogen bonding with other molecules of component. It is of great importance to perform a study on the mixtures involving alkanediols over the entire range of composition and temperature in order to see the effects of various parameters on the extension of hydrogen bonding of alkanediol mixtures which is influenced by some properties such as temperature, composition and structures. Moreover, alkanediols play an important role in industry that below we note to some of their properties and applications. 1,2-Ethanediol (ethylene glycol), 1,2-propanediol and 1,3–propanediol are members of alkanediol family which are produced by the vapor-phase oxidation of ethylene, by the hydration of propylene oxide derived from propylene by chlorohydrins process, and by the catalytic solution phase hydration of acrolein followed by reduction [12], respectively. Also, in recent years, many researchers have been attempting to find commercial method to produce 1,2-propanediol and 1,3-propanediol from biodiesel [13], [14]. The major applications of the investigated alkanediols in industry are as antifreezer, coolant, plasticizer, humectant, solvent and extractor in automotive, and also in tobacco, cosmetic, and petroleum industry [15], [16], [17]. Molecular structures of the studied alkanediols are shown in Fig. 1.

Many researchers have studied various binary solutions consisting 1,2-ethanediol, 1,2-propanediol, and 1,3-propanediol. George et al. [18] reported the density and dynamic viscosity, as well as speeds of sound and relative permittivities of water with alkanediols within the temperature range (298.15 to 338.15) K. Edward et al. [19] determined acoustic and thermodynamic properties of 1,2-propanediol and 1,3-propanediol at different temperatures and pressures. The Li group studied the density and excess properties of 1,4-butanediol with 1,2-ethanediol, 1,2-propanediol and 1,3-propanediol over the whole composition range [20]. Density and excess molar volume of Glycerol with 1,2-propanediol or 1,3-propanediol at various temperatures were studied by Li et al. [21]. Yang et al. [15] and Kapadi et al. [22] measured the density, viscosity and excess properties of the water with 1,2-ethanediol and 1,2-propanediol at different temperatures, respectively. Doghaei and coworkers investigated the density, viscosity, and volumetric properties of some alkanediols and alcohols binary mixtures at different temperatures [23]. Moreover, Survey of the more literature shows that many thermophysical properties of the other alkanediol mixtures have been reported [24], [25], [26], [27], [28], [29].

In this work, the experimental values of densities ρ and viscosities η of 1,2-ethanediol + 1,2-propanediol, + 1,3-propanediol and 1,2-propnediol + 1,3-propanediol binary mixtures over the entire concentration range and temperatures of (298.15 to 308.15) K were reported. The excess molar volume VmE, coefficient of thermal expansion αp, excess coefficient of thermal expansion αpE, isothermal coefficient of excess molar enthalpy HmEPT,x, partial molar volume V¯m,i, partial molar volume at infinite dilution V¯i, apparent molar volume Vφi, excess viscosity ηE and excess Gibbs energy of activation ΔG*E were calculated from the obtained experimental results. The values of VmE have been fitted to the Redlich–Kister polynomial equation to drive binary coefficients and estimate the standard deviation between experimental and calculated results. Moreover, Prigogine–Flory–Patterson (PFP) model has been applied to VmE for the present binary mixtures and the calculated PFP excess molar volumes were analyzed to determine the different parameters affecting VmE values. Furthermore, the viscosities of these binary mixtures were calculated theoretically using various semi-empirical relations such as Grunberg–Nissan, Tamura–Kurata, Hind–Ubbelohde and Katti–Chaudhary and also the obtained results were compared with the experimental data.

Section snippets

Materialsvd

1,3-propanediol was purchased from Merck. 1,2-Ethanediol and 1,2-propanediol were supplied by Fluka. The purity of these chemicals was higher than 98% and their water content was around 0.2%. All these compounds were used without further purification and the purities were verified by comparing the measured normal boiling points. The water content of the chemicals was measured with Kyoto mks-210 Karl Fischer instrument in order to be aware of the nature of impurities and their effect on

Results and discussion

The experimental values of densities along with their corresponding volumetric properties including coefficient of thermal expansion αp, excess molar volume VmE, partial molar volume V¯m,i, and apparent molar volume Vφi for binary mixtures of ethylene glycol + 1,2-propanediol or 1,3-propanediol and 1,2-propanediol + 1,3-propanediol as a function of mole fraction at temperatures 298.15, 308.15, and 308.15 K are reported in Table 3, Table 4, Table 5, respectively.

Density is the physical character of

Conclusion

This paper reports experimental data of densities and viscosities of 1,2-ethanediol + 1,2-propanediol or 1,3-propanediol and 1,2-propanediol + 1,3-propanediol binary mixtures over the whole composition range at temperatures (298.15 to 308.15) K with 5 K interval. The values of the pure components generally agree with the available literature data. The experimental values were used to compute the volumetric and viscosimetric parameters including the excess molar volume VmE, partial molar volume V¯m,i

Acknowledgment

The financial support from University of Mazandaran is gratefully acknowledged.

References (87)

  • E. Zorebski et al.

    Thermodynamic and transport properties of (1,2-ethanediol + 1-nonanol) at temperatures from (298.15 to 313.15) K

    J. Chem. Thermodyn.

    (2009)
  • N. Zhang et al.

    Excess molar volumes and viscosities of poly(ethylene glycol) 300 + water at different temperatures

    Fluid Phase Equilib.

    (2012)
  • A. Valtz et al.

    Liquid densities and excess molar volumes for water + diethylene glycolamine, and water, methanol, ethanol, 1-propanol + triethylene glycol binary systems at atmospheric pressure and temperatures in the range of 283.15–363.15 K

    Fluid Phase Equilib.

    (2004)
  • Z. Guo et al.

    Liquid viscosities, excess properties, and viscous flow thermodynamics of triethylene glycol + water mixtures at T = (298.15, 303.15, 308.15, 313.15, and 318.15) K

    J. Mol. Liq.

    (2012)
  • U. Doman'ska et al.

    Measurements of the density and viscosity of binary mixtures of (hyper-branched polymer, B-H2004 + 1-butanol, or 1-hexanol, or 1-octanol, or methyl tert-butyl ether)

    J. Chem. Thermodyn.

    (2010)
  • A.J. Treszczanowicz et al.

    Excess volumes of alkanol + alkane binary systems in terms of an association model with a Flory contribution term

    Fluid Phase Equilib.

    (1985)
  • R.B. Torres et al.

    Volumetric properties of binary mixtures of ethers and acetonitrile: experimental results and application of the Prigogine–Flory–Patterson theory

    J. Chem. Thermodyn.

    (2008)
  • K. Khanlarzadeh et al.

    Application of ERAS-model and Prigogine–Flory–Patterson theory to excess molar volumes for ternary mixtures of (2-chlorobutane + butylacetate + isobutanol) at T = 298.15 K

    J. Chem. Thermodyn.

    (2011)
  • M.T. Zafarani-Moattar et al.

    Application of Prigogine–Flory–Patterson theory to excess molar volume and speed of sound of 1-n-butyl-3-methylimidazolium hexafluorophosphate or 1-n-butyl-3-methylimidazolium tetrafluoroborate in methanol and acetonitrile

    J. Chem. Thermodyn.

    (2006)
  • M.L.L. Paredes et al.

    Densities, sound velocities, and refractive indexes of (tetralin + n-decane) and thermodynamic modeling by Prigogine–Flory–Patterson model

    J. Chem. Thermodyn.

    (2012)
  • T. Singh et al.

    Volumetric behavior of 1-butyl-3-Methyl imidazolium hexafluorophosphate with ethylene glycol derivatives: application of Prigogine–Flory–Patterson theory

    J. Mol. Liq.

    (2010)
  • A. Hartono et al.

    Density, viscosity, and excess properties of aqueous solution of diethylenetriamine (DETA)

    J. Chem. Thermodyn.

    (2009)
  • S.S. Dhondge et al.

    Volumetric and viscometric study of aqueous binary mixtures of some glycol ethers at T = (275.15 and 283.15) K

    J. Chem. Thermodyn.

    (2011)
  • M.J. Iqbal et al.

    Thermodynamic study of three pharmacologically significant drugs: Density, viscosity, and refractive index measurements at different temperatures

    J. Chem. Thermodyn.

    (2009)
  • K. Rajagopal et al.

    Excess parameter studies on binary liquid mixtures of 2-methyl-2-propanol with aliphatic nitriles at different temperatures

    J. Mol. Liq.

    (2011)
  • K. Narayanaswamy et al.

    Volumetric properties, viscosities, refractive indices and surface tensions for (dimethylpropanolamine (DMPA) + water) mixtures from 298.15 K to 343.15 K

    Thermochim. Acta

    (2012)
  • M. Yasmin et al.

    Experimental and computational study on viscosity and optical dielectric constant of solutions of poly (ethylene glycol) 200

    J. Mol. Liq.

    (2011)
  • X. Li et al.

    Thermodynamic and spectroscopic studies for aqueous solution of ethylene glycol monomethyl ether

    J. Mol. Liq.

    (2011)
  • A.K. Nain

    Ultrasonic and viscometric studies of molecular interactions in binary mixtures of formamide with ethanol, 1-propanol, 1,2-ethanediol and 1,2-propanediol at different temperatures

    J. Mol. Liq.

    (2008)
  • D.M. Bajić et al.

    Densities, viscosities and refractive indices of the binary systems PEG200 + 1,2-propanediol,  + 1,3-propanediol and PEG400 + 1,2-propanediol, + 1,3-propanediol at 288. 15 to 333. 15 K and atmospheric pressure: measurements and modeling

    J. Chem. Thermodyn.

    (2013)
  • D.B.K. Kumar et al.

    Thermodynamic and spectroscopic study of molecular interactions in the binary liquid mixtures of N-methyl-2-pyrrolidone and some substituted benzenes at different temperatures

    J. Mol. Liq.

    (2012)
  • A. Awasthia et al.

    Intermolecular interactions in formamide + 2-alkoxyethanols: viscometric study

    Thermochim. Acta

    (2012)
  • M.R. Kula et al.
    (1982)
  • C.M. Kinart

    Intermolecular Interactions in liquid formamide-dioxane mixtures found by measuring their 1H NMR spectra, densities, viscosities and relative dielectric permittivities

    Pol. J. Chem.

    (1993)
  • G.D. Yadav et al.

    Hydrogenolysis of glycerol to 1,2-propanediol over nano-fibrous Ag-OMS-2 catalysts

    Ind. Eng. Chem. Res.

    (2012)
  • P. boonoun et al.

    Application of sulfonated carbon-based catalyst for reactive extraction of 1,3-propanediol from model fermentation mixture

    Ind. Eng. Chem. Res.

    (2010)
  • C. Yang et al.

    Excess molar volumes, viscosities, and heat capacities for the mixtures of ethylene glycol + water from 273.15 K to 353.15 K

    J. Chem. Eng. Data

    (2003)
  • G.O. Curme et al.

    Ethylene glycol, a contribution of chemistry to the automobile antifreeze problem

    Ind. Eng. Chem.

    (1925)
  • A.H. Mohammadi et al.

    Experimental data and prediction of dissociation condition for ethane and propane simple hydrate in the presence of methanol, ethanol, ethylene glycol, and triethylene glycol aqueous solutions

    J. Chem. Eng. Data

    (2008)
  • J. George et al.

    Dynamic viscosities, speeds of sound and relative permittivities for water + alkanediols (propane-1,2- and -1,3-diol and butane-1, 2-, -1,3-, -1,4-, and -2,3-diol) at different temperatures

    J. Chem. Eng. Data

    (2003)
  • E. Zorebski et al.

    Study of the acoustic and thermodynamic properties of 1, 2- and 1,3-propanediol by means of high-pressure speed of sound measurements at temperatures from (293 to 318) K and pressures up to 101 MPa

    J. Chem. Eng. Data

    (2008)
  • Q.S. Li et al.

    Densities and excess molar volumes for binary mixtures of 1,4-butanediol + 1, 2-propanediol, + 1,3-propanediol, and + ethane-1,2-diol from (293.15 to 328.15) K

    J. Chem. Eng. Data

    (2008)
  • Q.S. Li et al.

    Densities and excess molar volumes for binary glycerol + 1-propanol, + 2-propanol, + 1, 2-propanediol, and + 1,3-propanediol mixtures at different temperatures

    J. Chem. Eng. Data

    (2007)
  • Cited by (13)

    • Thermodynamic, transport and optical properties of formamide + 1,2-ethanediol, 1,3-propanediol and poly (ethylene glycol) 200 binary liquid mixtures

      2017, Journal of Molecular Liquids
      Citation Excerpt :

      Thus, it is favorable to use this amide to model and study of the phase behavior of association systems in pure and the mixtures with other polar systems [8,9]. Alkanediols are the class of strong associating components with the ability to form inter-intra molecular hydrogen bonding [10]. It is of great interest to perform a study on the mixtures involving FA + diols at different composition and temperature to see the effects of various parameters on the extension of system's H-bonding.

    • Isobaric phase equilibrium at local and sub-atmospheric pressures, molar refractivity, volumetric and viscometric properties of toluene (1) + para-tert-butyltoluene (2) at temperatures (293.15 to 308.15) K: Experimental studies and thermodynamic modeling

      2016, Journal of Molecular Liquids
      Citation Excerpt :

      The deviation (AAD = 1/n ∑|((ηexp - η cal)/ηexp)| * 100) between experimental and calculated viscosity has been represented in Table 13. It was observed from Table 13, Grunberg-Nissan interaction parameter G12 is negative indicating the dominance of dispersive forces on mixing the component [34]. Fig. 6 based on AAD, calculated for different relations, it can be concluded that the viscosity predicted from Grunberg-Nissan are in good agreement with experimental values at 293.15 and 303.15 K. Heric and Brewer expression shows least deviations from experimental values at 298.15 K and 308.15 K.

    • Excess properties and isobaric (vapor + liquid) equilibrium at sub-atmospheric pressures of binary (1,2-propanediol + 1,3-propanediol) system: Measurement and modelling

      2016, Journal of Chemical Thermodynamics
      Citation Excerpt :

      Knowledge of (vapor + liquid) equilibrium (VLE) data of 1,2-pd + 1,3-pd is necessary for separation of components from their mixtures. The study of thermophysical properties such as density and refractive index of mixtures helps in understanding not only the interactions between solute–solute, solvent–solvent and solute–solvent, but also accounts for packing effects which arise due to differences in sizes of molecules [3]. There is an enormous significance to investigate the VLE behavior and thermodynamic properties of alkanediol mixtures as it plays an important role in industries.

    View all citing articles on Scopus
    View full text