Next Article in Journal
Bimodal Fucoidan-Coated Zinc Oxide/Iron Oxide-Based Nanoparticles for the Imaging of Atherothrombosis
Next Article in Special Issue
Subcritical Fluid Chromatography at Sub-Ambient Temperatures for the Chiral Resolution of Ketamine Metabolites with Rapid-Onset Antidepressant Effects
Previous Article in Journal
Convenient Synthesis of 6,7,12,13-Tetrahydro-5H-Cyclohepta[2,1-b:3,4-b’]diindole Derivatives Mediated by Hypervalent Iodine (III) Reagent
Previous Article in Special Issue
Supercritical CO2 Fluid Extraction of Elaeagnus mollis Diels Seed Oil and Its Antioxidant Ability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cohesive Energy Densities Versus Internal Pressures of Near and Supercritical Fluids

Institute of Analytical Chemistry of the Czech Academy of Sciences, Veveří 97, 60200 Brno, Czech Republic
Molecules 2019, 24(5), 961; https://doi.org/10.3390/molecules24050961
Submission received: 11 February 2019 / Revised: 6 March 2019 / Accepted: 7 March 2019 / Published: 8 March 2019
(This article belongs to the Special Issue Supercritical Fluids and Green Chemistry)

Abstract

:
Over half a century ago, Wiehe and Bagley suggested that a product of the internal pressure and molar volume of a liquid measures the energy of nonspecific intermolecular interactions whereas the cohesive energy reflects the total energy of intermolecular interactions in the liquid. This conjecture, however, has never been considered in connection with near and supercritical fluids. In this contribution, the cohesive energy density, internal pressure and their ratios are calculated from high precision equations of state for eight important fluids including water. To secure conformity to the principle of corresponding states when comparing different fluids, the calculations are carried out along the line defined by equality between the reduced temperature and the reduced pressure of the fluid (Tr = Pr). The results provide additional illustration of the tunability of the solvent properties of water that stands apart from those of other near and supercritical fluids in common use. In addition, an overview is also presented of the derivatives of cohesive energy density, solubility parameter and internal pressure with respect to temperature, pressure and molar volume.

Graphical Abstract

1. Introduction

Among all single-component solvents used in supercritical fluid technology, water has been recognized as the solvent with the greatest degree of tunability of the solvating abilities through changes in operating temperature and pressure [1,2]. Because of unique features of water and its importance as a solvent, a wide range of correlations have been developed for the important properties of water including relative permittivity [3,4] and ion product [5,6]. The variations in relative permittivity and ion product have been widely used to discuss/interpret the variable solvent properties of water. For example, at ambient conditions of 298.15 K and 0.1 MPa, the relative permittivity and ion product of water are 78.4 and 1.01 × 10−14 [mol.kg−1]2, respectively, whereas, at supercritical conditions of 773.15 K and 30 MPa, the relative permittivity and ion product of water drop to mere 1.7 and 1.2 × 10−21 [mol.kg−1]2, respectively [4,5]. These values illustrate that the solvent properties of water can be tuned within wide limits by properly adjusting the operating temperature and pressure. When interpreting and quantifying the solvent polarities in a broader sense, solvent effects on the absorption spectra of suitable organic probe molecules have also been frequented [7].
Another line of thought when considering the solvent properties may be based on strictly thermodynamic (PvT) properties of fluids. Over half a century ago, Wiehe and Bagley [8] suggested that a product of the internal pressure and molar volume of a liquid measures the energy of nonspecific intermolecular interactions whereas the cohesive energy reflects the total energy of intermolecular interactions in the liquid. To illustrate the concept of Wiehe and Bagley, let us recall the definition of cohesive energy density c,
c = u 0 u v
where v is the fluid’s molar volume, u is the fluid’s molar internal energy and u0 is the molar internal energy of the fluid in an ideal gas state at the particular temperature. The numerator in equation (1), u0u, is called cohesive energy and, obviously, it is the energy needed to break all intermolecular interactions in the fluid, whatever their kind. In turn, the internal pressure Pint of a fluid is the isothermal derivative of the fluid’s internal energy with respect to the fluid’s volume at a constant temperature,
P int = ( u v ) T = T ( P T ) V P = T γ V P
where T is the temperature, P is the pressure, V is the volume and γV is the thermal pressure coefficient. Wiehe and Bagley [8] assumed that, upon an infinitesimal change of volume, only nonspecific interactions in the fluid are affected, with specific interactions (hydrogen bonds) remaining intact. Wiehe and Bagley applied their concept to common liquid solvents near ambient temperature and pressure and, to the best of the present author’s knowledge, extension of the concept to near- and supercritical fluids has never been considered although the validity of Equations (1) and (2) is not limited to the liquid state. It follows from the Wiehe and Bagley’s original concept that the ratio c/Pint may serve as an approximate measure of the strength of total intermolecular interactions relative to the strength of nonspecific interactions.
The purpose of the present contribution is to discuss the behavior of c, Pint and c/Pint of eight popular fluids in their near and supercritical states to illustrate the unique position of water from yet another angle of view. To secure conformity to the principle of corresponding states when comparing different fluids, cohesive energy density, internal pressure and their ratio are calculated for the eight fluids along the line defined by equality between the reduced temperature and the reduced pressure of the fluid (Tr = Pr). An important feature of this choice is that the Tr = Pr line does not interfere with the vapor–liquid coexistence curve of the fluid (except at the critical point itself, Tr = Pr = 1). In addition, an overview is also presented of the derivatives of cohesive energy density, solubility parameter [9] and internal pressure with respect to temperature, pressure and molar volume.
It should also be noted here that the physical dimension of both internal pressure Pint and cohesive energy density c is pressure; therefore, the ratio c/Pint is dimensionless. In fact, the unit J.cm-3, often used for the cohesive energy density, is the same as MPa.

2. Results

In this section, the calculated results are discussed in the sequence starting from the thermal pressure coefficient γV, and continuing through the internal pressure Pint and the cohesive energy density c to the ratio c/Pint.

2.1. Thermal Pressure Coefficient

The plot in Figure 1 shows the calculated values of the thermal pressure coefficient as this quantity is an important constituent of the internal pressure of the fluid (see Equation (2)). It is readily apparent that, along the Tr = Pr line, the thermal pressure coefficient of water behaves very differently from those of the other fluids. The “hump” in the curve for water is only partly paralleled, to a much smaller extent, in the curves for methanol and ethanol while being absent in the other fluids.

2.2. Internal Pressure

The patterns in γV in the individual fluids (Figure 1) translate themselves into the respective plots for the internal pressure via Equation (2). Figure 2 indicates that the different behavior of the internal pressure between water and the other fluids is even more pronounced than the difference in the thermal pressure coefficient shown in Figure 1. It should be noted that the normal boiling temperature of water (100 °C, 373.15 K) corresponds to Tr = 0.5767. The maximum in the curve for water in Figure 2 appears at a temperature around 200 °C. Again, the plots for methanol and ethanol also show upward-convex regions although far less significant as compared with water.

2.3. Cohesive Energy Density

Unlike the plots in Figure 1, Figure 2 and Figure 3 suggests that, in the plots of cohesive energy density, there is no qualitative difference between water and the other fluids. Note, however, that the cohesive energy density of water exceeds those of the other fluids even at slightly supercritical conditions (Tr > 1).

2.4. Ratio of Cohesive Energy Density to Internal Pressure

As shown in Figure 4, the ratio c/Pint of water behaves very differently from those of the other fluids although, in fact, the difference occurs outside the domain of “near- and supercritical fluid”. At low subcritical temperatures (Tr < 0.6), c/Pint values of water far exceed those of the other fluids. In accord with the original concept of Wiehe and Bagley [8], c/Pint values are also rather high in methanol and ethanol, most probably because of intermolecular hydrogen bonding. Along the Tr = Pr line, the drop in c/Pint of water with rising temperature is much steeper as compared with methanol and ethanol, and, at Tr ~ 0.61, c/Pint of water drops even below that of methanol. This finding may appear somewhat surprising given the fact that the cohesive energy of water (see Figure 3) exceeds those of the other fluids even at supercritical conditions. As the temperature increases above the critical (Tr > 1), c/Pint converges to unity in most fluids including water. In the frame of Wiehe and Bagley’s concept [8], this finding would suggest that specific intermolecular interactions are no longer important in the near- and supercritical fluid region. The different course seen in methanol is likely due to a technical issue in the equation of state rather than to a fundamental feature of intermolecular interactions in methanol.

3. Materials and Methods

Comparative calculations of thermodynamic properties of different pure fluids have often been based on high precision equation-of-state packages such as ThermoFluids [10] or NIST Refprop [11]. Here, the calculations for the eight fluids have been carried out using the ThermoFluids software package [10] with embedded original equations of state or data for acetone [12], carbon dioxide [13], ethanol [14], methanol [15], n-pentane [16], R23 (trifluoromethane, fluoroform) [17], R134a (1,1,1,2-tetrafluoroethane) [18] and water [19]. In the calculations, the step in reduced temperature and pressure was 0.05 except between Tr = Pr from 0.95 to 1.05 where the step was 0.01. The upper limit of Tr = Pr was 1.40 in all fluids. The default value of the lower limit was 0.45, and it was applied in acetone, methanol, n-pentane and water. In the other fluids, this lower limit would fall within the solid phase region so that higher values had to be applied: 0.50 in ethanol and R134a, 0.55 in R23 and 0.75 in carbon dioxide, respectively. Calculation of the cohesive energy requires the molar internal energy in an ideal gas state. In this calculation, pressure was set to 1 × 10−6 Pa in all fluids at all temperatures. The calculated results have been available in the Supplementary Materials. All plots of the calculated properties have been prepared with the KyPlot graphing software (http://www.kyenslab.com), and the calculated data have been smoothed with piecewise cubic polynomials.

4. Discussion

Calculations based on high precision equations of state suggest that, at low subcritical temperatures along the Tr = Pr line, the ratio of cohesive energy density to internal pressure in water is significantly higher than in the other solvents. From a purely thermodynamic point of view, the reason for this is embedded in specific behavior of the thermal pressure coefficient of water along the Tr = Pr line. Some features similar to but much less important than the properties of water can also be seen in methanol and ethanol. These findings indicate that variations of the solvent properties of methanol and ethanol with temperature and pressure are far less important as compared with water. The calculated values of the cohesive energy density/internal pressure ratio provide yet another illustration of unique position of water among other near- and supercritical fluid solvents.
The weakest point of the present treatment may reside in the modeling of internal pressure; despite the century-long efforts spent on this problem [20,21,22], it still does not appear to have been resolved conclusively as regards the modeling of internal pressure in a wide range of temperature and pressure [23,24].
The variations of cohesive energy density and internal pressure of a fluid with state variables (PvT) usually have not received much attention, although they are quite important and an insight into their physical background seems desirable. In the Appendix A, therefore, an overview is presented of the derivatives of cohesive energy density, solubility parameter and internal pressure with respect to temperature, pressure and molar volume of the fluid. The expressions have been derived from Equations (1) and (2) employing standard thermodynamic relationships, and a list of symbols has been compiled at the end of Appendix A. It should also be mentioned here that the quantities appearing in Equations (A1)–(A15) are accessible from high precision equation-of-state packages such as ThermoFluids [10] or NIST Refprop [11], either directly or numerically (such as the derivatives of αP, βT, γV, cP and cV).

5. Conclusions

The calculations suggest that the ratio of cohesive energy density and internal pressure of water varies within a much wider range as compared with the other fluids included in this study. However, a large part of c/Pint variations with temperature and pressure occurs at low subcritical temperatures, that is, outside the domain of near- and supercritical fluid. In a broader sense, therefore, applicability of Wiehe and Bagley’s original concept remains confined to subcritical liquids because, in the near- and supercritical region, the ratio of cohesive energy density and internal pressure approaches unity. Within the frame of Wiehe and Bagley’s concept, this finding would suggest that specific intermolecular interactions become unimportant in the near- and supercritical region.

Supplementary Materials

The following are available online: A MS Excel file containing numerical values of all calculated properties needed to construct the plots in Figures 1–4.

Funding

This contribution has been supported by The Czech Science Foundation (Project No. 19-00742S) and by The Czech Academy of Sciences (Institutional Research Plan RVO:68081715).

Acknowledgments

The author thanks the anonymous reviewers whose insightful comments helped him to improve the manuscript.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A

Derivatives of c and Pint with Respect to T, P and v

The derivatives of cohesive energy density c are given by
( c T ) v = c V 0 c V v
( c T ) P = c V 0 c V v α P ( P int + c )
( c P ) T = β T ( P int + c )
( c P ) v = c V 0 c V v γ V
( c v ) T = ( P int + c ) / v
( c v ) P = 1 v [ c V 0 c V v α P P int c ]
The temperature derivative of cohesive energy density along the vapor–liquid coexistence curve may be written as
( c T ) σ = c V 0 c V v [ β T ( P T ) σ α P ] ( P int + c )
with (∂P/∂T)σ being the slope of the vapor-liquid coexistence curve. With the proper values inserted, Equation (A7) is valid for both liquid and vapor phases.
Since the solubility parameter δ of a fluid has been defined as the square root of the cohesive energy density, the derivatives of the solubility parameter are readily obtained from the corresponding derivatives described by Equations (A1)–(A7). Generally,
δ x = 1 2 δ c x
where ∂c/∂x is any of the derivatives given by Equations (A1)–(A7).
The derivatives of the internal pressure may be described as follows,
( P int T ) v = T ( γ V T ) v = ( c V v ) T = T ( 2 P T 2 ) v
( P int T ) P = γ V + T ( γ V T ) P = γ V + ( c P v ) T
( P int P ) T = T ( γ V P ) T 1 = T [ α P + v ( α P v ) T ] 1
( P int P ) v = T ( γ V P ) v = T [ 1 T γ V { ( c P v ) T c P γ V ( γ V v ) T } α P v ( α P v ) T ]
( P int v ) T = T ( γ V v ) T + 1 v β T = 1 v β T [ T β T ( β T T ) v + 1 ]
( P int v ) P = γ V v α P + T ( γ V v ) P = 1 v β T + T ( γ V v ) P
The temperature derivative of internal pressure along the vapor–liquid coexistence curve is given by
( P int T ) σ = γ V + T ( γ V T ) P + [ T ( γ V P ) T 1 ] ( P T ) σ = γ V + ( c P v ) T { T [ α P + v ( α P v ) T ] + 1 } ( P T ) σ

List of Symbols

ccohesive energy density
cPmolar isobaric heat capacity
cVmolar isochoric heat capacity
cV0ideal-gas molar isochoric heat capacity
Ppressure
Pintinternal pressure [= (∂u/∂v)T]
Ttemperature
umolar internal energy
u0ideal-gas molar internal energy
Vvolume
vmolar volume

Greek Symbols

αPisobaric expansivity [= (1/v)(∂v/∂T)P]
βTisothermal compressibility [= −(1/v)(∂v/∂P)T]
γVthermal pressure coefficient [= (∂P/∂T)V]
δsolubility parameter [= c1/2]

References

  1. Harvey, A.H.; Friend, D.G. Physical properties of water. In Aqueous Systems at Elevated Temperatures and Pressures. Physical Chemistry in Water, Steam and Hydrothermal Solutions; Palmer, D.A., Fernández-Prini, R., Harvey, A.H., Eds.; Elsevier–Academic Press: London, UK, 2004; Chapter 1; pp. 1–27. [Google Scholar]
  2. Weingärtner, H.; Franck, E.U. Supercritical water as a solvent. Angew. Chem. Int. Ed. 2005, 44, 2672–2692. [Google Scholar] [CrossRef] [PubMed]
  3. Uematsu, M.; Franck, E.U. Static dielectric constant of water and steam. J. Phys. Chem. Ref. Data 1980, 9, 1291–1304. [Google Scholar] [CrossRef]
  4. Fernández, D.P.; Goodwin, A.R.H.; Lemmon, E.W.; Levelt Sengers, J.M.H.; Williams, R.C. A formulation for the static permittivity of water and steam at temperatures from 238 K to 873 K at pressures up to 1200 MPa, including derivatives and Debye–Hückel coefficients. J. Phys. Chem. Ref. Data 1997, 26, 1125–1166. [Google Scholar] [CrossRef]
  5. Marshall, W.L.; Franck, E.U. Ion product of water substance, 0–1000 °C, 1–10,000 bars. New international formulation and its background. J. Phys. Chem. Ref. Data 1981, 10, 295–304. [Google Scholar] [CrossRef]
  6. Bandura, A.V.; Lvov, S.N. The ionization constant of water over wide ranges of temperature and density. J. Phys. Chem. Ref. Data 2006, 35, 15–30. [Google Scholar] [CrossRef]
  7. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2003; Chapter 6; pp. 329–388. [Google Scholar]
  8. Wiehe, I.A.; Bagley, E.B. Estimation of dispersion and hydrogen bonding energies in liquids. AIChE J. 1967, 13, 836–838. [Google Scholar] [CrossRef]
  9. Prausnitz, J.M.; Lichtenthaler, R.N.; Gomes de Azevedo, E. Molecular Thermodynamics of Fluid-Phase Equilibria, 3rd ed.; Prentice Hall: Upper Saddle River, NJ, USA, 1999; Section 7.2; pp. 313–326. [Google Scholar]
  10. Wagner, W.; Overhoff, U. ThermoFluids. Interactive Software for the Calculation of Thermodynamic Properties for More Than 60 Pure Substances; Springer-Verlag: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  11. NIST Reference Fluid Thermodynamic and Transport Properties Database (REFPROP). Available online: https://www.nist.gov/srd/refprop (accessed on 7 February 2019).
  12. Lemmon, E.W.; Span, R. Short fundamental equations of state for 20 industrial fluids. J. Chem. Eng. Data 2006, 51, 785–850. [Google Scholar] [CrossRef]
  13. Span, R.; Wagner, W. A new equation of state for carbon dioxide covering the fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa. J. Phys. Chem. Ref. Data 1996, 25, 1509–1596. [Google Scholar] [CrossRef]
  14. Dillon, H.E.; Penoncello, S.G. A fundamental equation for calculation of the thermodynamic properties of ethanol. Int. J. Thermophys. 2004, 25, 321–335. [Google Scholar] [CrossRef]
  15. de Reuck, K.M.; Craven, R.J.B. International Thermodynamic Tables of the Fluid State–Vol. 12: Methanol; Blackwell Scientific Publications: Oxford, UK, 1993. [Google Scholar]
  16. Span, R.; Wagner, W. Equations of state for technical applications. II. Results for nonpolar fluids. Int. J. Thermophys. 2003, 24, 41–109. [Google Scholar] [CrossRef]
  17. Penoncello, S.G.; Shan, Z.; Jacobsen, R.T. A fundamental equation for the calculation of the thermodynamic properties of trifluoromethane (R-23). ASHRAE Trans. 2000, 106, 739–756. [Google Scholar]
  18. Tillner-Roth, R.; Baehr, H.D. An international standard equation of state for the thermodynamic properties of 1,1,1,2-tetrafluoroethane (HFC-134a) for temperatures from 170 K to 455 K at pressures up to 70 MPa. J. Phys. Chem. Ref. Data 1994, 23, 657–729. [Google Scholar] [CrossRef]
  19. Wagner, W.; Pruss, A. The IAPWS formulation 1995 for the thermodynamic properties of ordinary water substance for general and scientific use. J. Phys. Chem. Ref. Data 2002, 31, 387–535. [Google Scholar] [CrossRef]
  20. Lewis, W.C.M. Internal, molecular, or intrinsic pressure—A survey of the various expressions proposed for its determination. Trans. Faraday Soc. 1911, 7, 94–115. [Google Scholar] [CrossRef]
  21. Hildebrand, J.H. Solubility III. Relative values of internal pressures and their practical application. J. Am. Chem. Soc. 1919, 41, 1067–1080. [Google Scholar] [CrossRef]
  22. Dack, M.R.J. The importance of solvent internal pressure and cohesion to solution phenomena. Chem. Soc. Rev. 1975, 4, 211–229. [Google Scholar] [CrossRef]
  23. Marcus, Y. Internal pressure of liquids and solutions. Chem. Rev. 2013, 113, 6536–6551. [Google Scholar] [CrossRef] [PubMed]
  24. Chorążewski, M.; Postnikov, E.B.; Oster, K.; Polishuk, I. Thermodynamic properties of 1,2-dichloroethane and 1,2-dibromoethane under elevated pressures: Experimental results and predictions of a novel DIPPR-based version of FT-EoS, PC-SAFT, and CP-PC-SAFT. Ind. Eng. Chem. Res. 2015, 54, 9645–9656. [Google Scholar] [CrossRef]
Figure 1. Thermal pressure coefficients γV of the selected fluids.
Figure 1. Thermal pressure coefficients γV of the selected fluids.
Molecules 24 00961 g001
Figure 2. Internal pressures Pint of the selected fluids.
Figure 2. Internal pressures Pint of the selected fluids.
Molecules 24 00961 g002
Figure 3. Cohesive energy densities c of the selected fluids.
Figure 3. Cohesive energy densities c of the selected fluids.
Molecules 24 00961 g003
Figure 4. Cohesive energy density/internal pressure ratios c/Pint of the selected fluids.
Figure 4. Cohesive energy density/internal pressure ratios c/Pint of the selected fluids.
Molecules 24 00961 g004

Share and Cite

MDPI and ACS Style

Roth, M. Cohesive Energy Densities Versus Internal Pressures of Near and Supercritical Fluids. Molecules 2019, 24, 961. https://doi.org/10.3390/molecules24050961

AMA Style

Roth M. Cohesive Energy Densities Versus Internal Pressures of Near and Supercritical Fluids. Molecules. 2019; 24(5):961. https://doi.org/10.3390/molecules24050961

Chicago/Turabian Style

Roth, Michal. 2019. "Cohesive Energy Densities Versus Internal Pressures of Near and Supercritical Fluids" Molecules 24, no. 5: 961. https://doi.org/10.3390/molecules24050961

Article Metrics

Back to TopTop