Elsevier

Thermochimica Acta

Volume 570, 20 October 2013, Pages 16-26
Thermochimica Acta

Volume properties of liquid mixture of water + glycerol over the temperature range from 278.15 to 348.15 K at atmospheric pressure

https://doi.org/10.1016/j.tca.2013.07.012Get rights and content

Highlights

  • Temperature-dependent densities of water + glycerol mixtures were measured.

  • Excess molar volumes and partial molar volumes of components were calculated.

  • Coefficients of isobaric expansion for mixture and components were discussed.

Abstract

Densities of {water (1) + glycerol (2)} binary mixture have been measured with vibrational densimeter over the whole concentration range within the temperature interval 278.15–348.15 K. Excess molar volumes and molar isobaric expansions of the mixture, apparent and partial molar volumes, as well partial molar isobaric expansions of water and glycerol in the mixture, limiting partial molar volumes and limiting partial molar isobaric expansions of water and glycerol have been calculated. It was shown that the mixture densities increased on the glycerol molar fraction rising at every temperature studied, and at low glycerol concentrations this increase was more pronounced. Magnitudes of excess molar volumes of the mixture were negative at all temperatures. But the deviation from ideality decreased with temperature growth. Molar isobaric expansions of the mixture increased almost linearly when glycerol concentration went up. Concentration dependences of partial molar isobaric expansions of water and glycerol were characterized by extremes occurrence.

Introduction

The solvents forming spatial H-bond network and their mixtures possess some particular properties such as relatively large free volume, small isobaric expansion and isothermal compressions, high viscosity and its strong dependence on temperature, unlimited miscibility with other H-bonded solvents, and so on. Triols, along with water and diols, belong to the solvents with developed spatial network of hydrogen bonds. The simplest representative of triols is glycerol, being studied extensively [1], [2], [3], [4], [5], [6], [7], [8], [9], [10], [11]. Earlier we have studied the volume properties of water–ethylene glycol mixture [12] and in this connection the volume properties of water–glycerol mixture arouse interest.

In practice water + glycerol mixture is widely used as cryoprotectors for preservation of biomaterials used for damping of intracellular ice formation and damage of biological tissues [13], [14], [15]. In nature marine algae, salt-tolerant plants, insects and fish living at low temperatures contain glycerol as a common cellular component. The stability of biological molecules is explained precisely by glycerol presence [16], [17]. Thus, it is applied as general cryoprotector which increases liquids ability to osmosis and preserves them from ice formation [18].

Glycerol (Gly, 1,2,3-propanetriol) has the following physical properties: ɛ = 41.14, μ = 0.28 D, η = 1.45 Pa s at 293 K; Tmelting  291.3 K; Tboiling  563 K (with decomposition) [19], [20], [21]. In Gly molecule there are three hydrophilic alcoholic groups forming intra- and intermolecular hydrogen bonds. These three alcohol groups are considered to be responsible for Gly unlimited solubility in water. Liquid glycerol demonstrates anomalous temperature behavior of viscosity coefficient [22], [23] and dielectric relaxation times [24], [25]. Both these phenomena are determined, according to literature, by H-bond network formation [26]. Physical properties of glycerol depend on external parameters and H-bonds number varies with temperature and pressure [27], [28]. Glycerol can exist as supercooled liquid that makes its crystallization possible just with special technique [5], [28]. Mere temperature decreasing (down to 185 K) brings usually to glassy state formation [29]. Glycerol molecule is very flexible and can theoretically form, according to results of computer simulation, 126 conformers [4], [30], [31]. If in crystalline state Gly molecule is as αα conformer [5], then the data on quality composition of conformers in other aggregate states are rather inconsistent. In [6] it is asserted that glassy glycerol forms also one but another conformation – βγ. In liquid state the presence of two dominating conformers – αα and αγ – has been confirmed experimentally. However the possibility of another conformers existing is not ruled out [1], [2], [7], [31].

Last years water–glycerol system was studied intensively by various methods such as molecular dynamic [4], [8], [26], [31], [32], [33], [34], [35], [36], [37], [38], [39], diverse thermodynamic measurements [40], [41], [42], NMR [43], infrared [44] and Raman spectroscopy [43], [45], [46]. The system volume properties have been studied also. Before, the density of water–glycerol system, due to its technological importance, has been measured for the first time as far back as 1884 [47]. In the reference-book [48] published in 1928 all results obtained were summarized and the mixture densities were performed at 15, 15.5, 20 and 25 °C (±0.1 °C). But surprisingly till now the majority of density precise measurements of water–glycerol mixture has been carried out either only at one temperature (293.15 or 298.15) or within narrow temperature interval. So these data do not allow calculating the mixture expansion with high accuracy. Besides the works dealing with supercooled glycerol are rather scarce.

The present work continues our earlier investigations of volume properties of binary mixtures with different nature of intermolecular interaction depending on composition, temperature, and pressure [49], [50], [51], [52], [53].

Section snippets

Experimental part

In the work the solvents of the highest available purity were used. The initial glycerol (stated purity 99.5%, “Khimreactive”) was purified by double distillation according to reference [21] and was kept under vacuum. The water content was determined by K. Fischer's method and did not exceed 0.02 wt.% (or 4 × 10−5 molar fraction). For solutions preparation the deionized freshly prepared bidistillate was used.

The mixtures were prepared by gravimetric method with accuracy 1 × 10−3 g from degasified

Calculations

Excess molar volumes of the mixture VmE:VmE=Vmx1V1ox2V2owhere Vm is the mixture molar volume, V1o, x1, and V2o, x2 are molar volumes of pure components and their molar fractions (indexes 1 and 2 refer to water and glycerol, accordingly). VmE values were found directly from experimental data by equation:VmE=x1M11ρ1ρ1+x2M21ρ1ρ2where M1, ρ1, and M2, ρ2 are molar weights and densities of water and glycerol, accordingly; ρ is the mixture density. Uncertainty of excess molar volumes determination

Discussion

Intermolecular interactions in the mixed solvent of {water (1) + glycerol (2)} are mainly determined by hydrogen bonds between hydroxyl groups. Variation of the mixture volume, first of all, depends on alteration of H-bonds energy and their quantity. Moreover the mixture volume reflects conformational equilibrium of glycerol molecules. It is known that in liquid phase several conformations of glycerol molecules can exist simultaneously [1], [2], [3], [4], [30], [31], but depending on external

Conclusions

Volume characteristics of liquid {water(1) + glycerol (2)}mixture reveal that in the mixture complex processes run upon temperature and composition variation. They are caused by rearrangement of H-bond components participating into intermolecular interaction that also leads to conformational variations.

The mixture formation is attended by more compact packing formation. On rising the temperature the absolute value of excess molar volume at a certain composition decreases.

Partial molar volumes of

Acknowledgment

This work was supported by the Russian Foundation for Basic Research (project 12-03-97525-r_centre_a) and grant of the President of the Russian Federation (No. MK-1288.2013.3).

References (83)

  • 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)
  • Selected Values of Properties of Chemical Compounds, Data Project

    (1980)
  • L. Verhoeye et al.

    Vapor–liquid equilibrium of the system 2-propanol-water-l,2,3-propanetriol at 760 mm of Hg

    J. Chem. Eng. Data

    (1969)
  • M. Sethu Raman et al.

    Ultrasonic and DFT study of intermolecular association through hydrogen bonding in aqueous solutions of glycerol

    J. Mol. Liq.

    (2008)
  • G.I. Egorov et al.

    The p-Vm-T-x properties of water–acetone mixtures over the temperature range 298–323 K and pressures from 1 to 1000 bar. I. Experimental results and partial molar volumes

    Russ. J. Phys. Chem. A

    (1996)
  • J.-Y. Huot et al.

    A comprehensive thermodynamic investigation of water–ethylene glycol mixtures at 5, 25, and 45 ̊C

    J. Solut. Chem.

    (1988)
  • J.J. Towey et al.

    The structure of glycerol in the liquid state: a neutron diffraction study

    Phys. Chem. Chem. Phys.

    (2011)
  • D.C. Champeney et al.

    Structural studies of liquid d-glycerol by neutron diffraction

    Mol. Phys.

    (1986)
  • R. Chelli et al.

    Density functional calculation of structural and vibrational properties of glycerol

    J. Phys. Chem. A

    (2000)
  • R. Chelli et al.

    Glycerol condensed phases. Part I. A molecular dynamics study

    Phys. Chem. Chem. Phys.

    (1999)
  • H. van Koningsveld

    The crystal structure of glycerol and its conformation

    Rec. Trav. Chim.

    (1968)
  • J. Dawidowski et al.

    Coherent neutron scattering response from glassy glycerol

    Phys. Rev. E

    (1996)
  • M. Garawia et al.

    Structural studies of liquid d-glycerol II. Molecular conformation and long range correlations

    Mol. Phys.

    (1987)
  • J.L. Dashnau et al.

    Hydrogen bonding and the cryoprotective properties of glycerol/water mixtures

    J. Phys. Chem. B

    (2006)
  • P. Jain et al.

    Molecular dynamics in supercooled glycerol: results from 13C NMR spectroscopy

    J. Chem. Phys.

    (2009)
  • Y. Marcus

    Some thermodynamic and structural aspects of mixtures of glycerol with water

    Phys. Chem. Chem. Phys.

    (2000)
  • P. Westh et al.

    Intermolecular interactions in ternary glycerol–sample–H2O: towards understanding the Hofmeister series (V)

    J. Solut. Chem.

    (2011)
  • G.I. Egorov et al.

    Volumetric properties of the water–ethylene glycol mixtures in the temperature range 278–333.15 K at atmospheric pressure

    Russ. J. Gen. Chem.

    (2010)
  • A. Benamotz et al.

    The role of glycerol in the osmotic regulation of the halophilic alga Dunaliella parva

    Plant Physiol.

    (1973)
  • K. Gekko et al.

    Mechanism of protein stabilization by glycerol: preferential hydration in glycerol–water mixtures?

    Biochemistry

    (1981)
  • J.A. Raymond

    Glycerol is a colligative antifreeze in some northern fishes

    J. Exp. Zool.

    (1992)
  • P.H. Yancey et al.

    Living with water stress: evolution of osmolyte systems

    Science

    (1982)
  • S. Garcia-Manyes et al.

    Osmolyte-induced separation of the mechanical folding phases of ubiquitin

    Proc. Natl. Acad. Sci. U.S.A.

    (2009)
  • P. Mazur

    Cryobiology: the freezing of biological systems

    Science

    (1970)
  • V.A. Rabinovich et al.
    (1991)
  • F. Weissberger et al.

    Organic Solvents: Physical Properties and Methods of Purification

    (1955)
  • R.L. Cook et al.

    Pressure and temperature dependent viscosity of two glass forming liquids: glycerol and dibutyl phthalate

    J. Chem. Phys.

    (1994)
  • D.B. Davies et al.

    Viscoelastic behaviour of some viscous liquids

    J. Chem. Soc. Faraday Trans.

    (1973)
  • D.C. Champeney et al.

    Ionic mobility and dielectric relaxation in supercooled liquid KCl–glycerol solutions

    Mol. Phys.

    (1984)
  • N. Menon et al.

    Evidence for a divergent susceptibility at the glass transition

    Phys. Rev. Lett.

    (1995)
  • Cited by (0)

    View full text