You are on page 1of 15

entropy

Article
Thermodynamic Fluid Equations-of-State
Leslie V. Woodcock
Department of Physics, University of Algarve, 8005-139 Faro, Portugal; lvwoodcock@ualg.pt

Received: 24 October 2017; Accepted: 30 December 2017; Published: 4 January 2018

Abstract: As experimental measurements of thermodynamic properties have improved in accuracy, to


five or six figures, over the decades, cubic equations that are widely used for modern thermodynamic
fluid property data banks require ever-increasing numbers of terms with more fitted parameters.
Functional forms with continuity for Gibbs density surface ρ(p,T) which accommodate a critical-point
singularity are fundamentally inappropriate in the vicinity of the critical temperature (Tc ) and
pressure (pc ) and in the supercritical density mid-range between gas- and liquid-like states.
A mesophase, confined within percolation transition loci that bound the gas- and liquid-state by
third-order discontinuities in derivatives of the Gibbs energy, has been identified. There is no
critical-point singularity at Tc on Gibbs density surface and no continuity of gas and liquid. When
appropriate functional forms are used for each state separately, we find that the mesophase pressure
functions are linear. The negative and positive deviations, for both gas and liquid states, on either
side of the mesophase, are accurately represented by three or four-term virial expansions. All gaseous
states require only known virial coefficients, and physical constants belonging to the fluid, i.e., Boyle
temperature (TB ), critical temperature (Tc ), critical pressure (pc ) and coexisting densities of gas
(ρcG ) and liquid (ρcL ) along the critical isotherm. A notable finding for simple fluids is that for all
gaseous states below TB , the contribution of the fourth virial term is negligible within experimental
uncertainty. Use may be made of a symmetry between gas and liquid states in the state function
rigidity (dp/dρ)T to specify lower-order liquid-state coefficients. Preliminary results for selected
isotherms and isochores are presented for the exemplary fluids, CO2 , argon, water and SF6 , with
focus on the supercritical mesophase and critical region.

Keywords: equation-of-state; liquid-gas criticality; carbon dioxide; argon; water; SF6 ; virial coefficients

1. Introduction
There is a long history and an extensive literature of cubic equations-of-state, going back 140 years,
from van der Waals renowned two-term equation for Andrew’s original p-V-T data on carbon
dioxide [1], to current research and compilations with hundreds of terms and parameters. As the
thermodynamic experimental measurements have improved in accuracy, ever more complex equations
have evolved [2]. These equations are used to represent experimental thermodynamic properties for
modern data banks such as the NIST compilation for CO2 [2,3]. Over recent decades, ever-increasing
numbers of terms and more fitted parameters are required. The Span–Wagner equation for CO2 [2],
for example, contains hundreds of terms, each with many fitted parameters, and various adjustable
fractional exponents. As even more accurate data become available from future research, these
functional forms will need yet more adjustable parameters.
The basic reason of this inconvenient modeling practice is the van der Waals gas-liquid continuity
hypothesis [1]. Continuous cubic functional forms are inappropriate in the vicinity of Tc and in the
supercritical mid-range between gas and liquid phases. A mesophase, confined within percolation
loci that bound the existence of gas and liquid phases by higher-order discontinuities, has been
identified [4–6]. A simple numerical differentiation of NIST equations-of-state [7] can demonstrate the
supercritical mesophase and observe the phase bounds, along any isotherm, of any fluid (e.g., CO2

Entropy 2018, 20, 22; doi:10.3390/e20010022 www.mdpi.com/journal/entropy


Entropy 2018, 20, 22 2 of 15

Entropy 2018, 20, 22 2 of 15

Figure 1) for any of the 200 fluids in the NIST Thermophysical Property data bank. These boundaries
any of the 200 fluids in the NIST Thermophysical Property data bank. These boundaries have been
have smoothed
been smoothed
over byover by the equations-of-state
the equations-of-state used to parameterize
used to parameterize the originaldata.
the original experimental experimental data.

CO2 supercritical isotherms


40 750 700
600
35 500
450
30
400 340
25 375 330
p 20 350
320
(MPa)
15 310
10
305
5

0
0 5 10 15 20

density (mol/L)
Figure 1. Supercritical T(K) isotherms for CO2: the green dashed line is the percolation loci PB; the blue
Figure 1. Supercritical T(K) isotherms for CO2 : the green dashed line is the percolation loci PB; the blue
dashed line is the percolation loci PA; the green and blue solid circles are the maximum experimentally
dashed line is the percolation loci PA; the green and blue solid circles are the maximum experimentally
observable coexisting gas density, and minimum coexisting liquid density along Tc i.e., 305 K (purple).
observable coexisting gas density, and minimum coexisting liquid density along Tc i.e., 305 K (purple).
Within the uncertainties, p(ρ,T) is a linear function of ρ in the mesophase region. The origin of
the linearity
Within is the colloidalp(ρ,T)
the uncertainties, nature isofathe supercritical
linear functionmesophase
of ρ in the[6,8] with a linear
mesophase combination
region. The origin ruleof the
for thermodynamic state functions, similar to subcritical lever rule for the inhomogeneous
linearity is the colloidal nature of the supercritical mesophase [6,8] with a linear combination rule for 2-phase
region. When appropriate functional forms are used for gaseous and liquid states separately, we find
thermodynamic state functions, similar to subcritical lever rule for the inhomogeneous 2-phase region.
that all other isothermal thermodynamic state functions in the mesophase are linear functions of
When appropriate functional forms are used for gaseous and liquid states separately, we find that all
density. In the case of pressure, the negative and positive adjacent deviations for liquid and gas on
othereither
isothermal
side are thermodynamic state functions inrequire
quadratic. The parameterizations the mesophase
only physicalare linear functions
constants belongingof density.
to the In
the case of pressure, the negative and positive adjacent deviations for liquid
specific fluid, Boyle temperature (TB), critical temperature (TC), and coexisting densities along theand gas on either side are
quadratic.
criticalThe parameterizations
isotherm, and known virial require only physical
coefficients b2(T) andconstants
b3(T) etc.belonging
A remarkable to the specific
finding fluid,
is that forBoyle
temperature (TB ), critical
the gas phases below T temperature
B, the coefficient(TC ),b4 and coexisting
is effectively densities
zero within the along the critical
experimental isotherm, and
uncertainty.
known There is rigidity
virial symmetry
coefficients b2 (T) between
and b3 (T) gasetc.
andAliquid phases finding
remarkable on eitherisside
thatoffor
thethe
mesophase
gas phasesreported
below TB ,
previously [9], that relates the lower coefficients of the liquid equation to
the coefficient b4 is effectively zero within the experimental uncertainty. There is rigidity symmetryproperties of the gas in the
vicinity of the percolation loci.
between gas and liquid phases on either side of the mesophase reported previously [9], that relates the
The objective of this research is to investigate the extent that these findings can be used to describe
lower coefficients of the liquid equation to properties of the gas in the vicinity of the percolation loci.
experimental thermodynamic properties of a pure fluid over the whole range of equilibrium existence.
The objective of this research is to investigate the extent that these findings can be used to describe
Here, we investigate equations-of-state for gas, liquid and mesophase separately, each of which, in
experimental
an initial thermodynamic
approximation, require properties
only of a pure fluid
measurable over the
physical whole specific
constants range oftoequilibrium existence.
particular fluids.
Here,Comparisons
we investigate with equations-of-state
experimental data are formade
gas, via
liquid and mesophase
the NIST thermo-physical separately,
databankeach of which,
[3]. NIST
in antabulations
initial approximation, require only
reproduce experimental measurable
data with 100% physical
accuracy,constants
and therefore,specific to particular
provide a foremost fluids.
Comparisons
criteria forwith experimental
testing the alternative data are madetovia
description vanthederNIST
Waalsthermo-physical
underlying science databank
of critical[3].
andNIST
supercritical
tabulations behavior
reproduce of fluids as data
experimental described
with in references
100% [4–9].
accuracy, and therefore, provide a foremost criteria
for testing the alternative description to van der Waals underlying science of critical and supercritical
behavior of fluids as described in references [4–9].
Entropy 2018, 20, 22 3 of 15

2. Rigidity and Fluid-State Bounds


Rigidity, (ω)T , is the work required to isothermally and reversibly increase the density of a fluid.
With dimensions of a molar energy, this simple state function relates directly to the change in Gibbs
energy (G) with density at constant T
Entropy 2018, 20, 22 3 of 15

ωT = (dp/dρ)T = ρ(dG/dρ)T
2. Rigidity and Fluid-State Bounds (1)

Rigidity,that
The inequalities , is the work gas
(ω)Tdistinguish required
fromtoliquid
isothermally
are: and reversibly increase the density of a fluid.
With dimensions of a molar energy, this simple state function relates directly to the change in Gibbs
energy (G) with density at constant T 2
GAS ρ < ρPB (d ω/dρ)T < 0 (2)
ωT = (dp/dρ)T = ρ(dG/dρ)T (1)
LIQUID ρ>ρ (d2 ω/dρ)T > 0 (3)
The inequalities that distinguish gas from liquidPA
are:
and for the mesophase GAS ρ < ρPB (d2ω/dρ)T < 0 (2)
2
MESO ρPB < ρρ<> ρPA
LIQUID PA (d(d T > 0T = 0
ω/dρ)
2ω/dρ) (3) (4)
and for the mesophase
The rigidity isotherms shown in Figure 2 for CO2 have been obtained from NIST Thermophysical
databank [3]. Inequalities in the MESO ρPB < of
derivatives ρ <ω ρPA (d2ω/dρ)T = 0
T (Equations (2) and (3)) can thermodynamically
(4)
define theThe percolation loci [9].shown
rigidity isotherms It follows that2 at
in Figure forlow
CO2density
have been the percolation
obtained loci must
from NIST approach the
Thermophysical
Boyledatabank
temperature (TB ) by its in
[3]. Inequalities definition, whichofisωobtained
the derivatives from
T (Equations (2)the
andrigidity
(3)) canintersection interpolated
thermodynamically
at zerodefine the percolation
density. For CO2 the locilimiting
[9]. It follows that(RT
rigidity at Blow density
: where R isthe percolation
the loci must
gas constant) alongapproach the
the percolation
Boyle temperature (T ) by its definition, which is obtained from the rigidity
loci is 6.03 kJ/mol and TB is 725 K, i.e., the temperature above which the 2nd-virial coefficient
B intersection interpolated at (b2 )
zero density. For CO 2 the limiting rigidity (RTB: where R is the gas constant) along the percolation loci
is positive and below which it is negative; at TB , p = ρkT and b2 = 0. In the pressure-density plane,
is 6.03 kJ/mol and TB is 725 K, i.e., the temperature above which the 2nd-virial coefficient (b2) is positive
the percolation loci exhibit maxima as density decreases and approach zero, i.e., the ideal gas limit at
and below which it is negative; at TB, p = ρkT and b2 = 0. In the pressure-density plane, the percolation
low pressures and densities. The same behavior is seen for other atomic (e.g., argon) and molecular
loci exhibit maxima as density decreases and approach zero, i.e., the ideal gas limit at low pressures and
(e.g., densities.
water) fluids.
The same behavior is seen for other atomic (e.g., argon) and molecular (e.g., water) fluids.

Figure 2. Rigidity of fluid phases of CO2 from NIST thermo-physical tables: the loci of gas and
Figure 2. Rigidity of fluid phases of CO2 from NIST thermo-physical tables: the loci of gas and
liquid-phase bounds are green (percolation line PB) and blue (percolation line PA) respectively; red
liquid-phase bounds areisotherms,
lines are supercritical green (percolation line
blue lines are PB) andisotherms;
subcritical blue (percolation
the purpleline PA) respectively;
isotherm is Tc (305 K): red
lines the
are Boyle
supercritical isotherms,
temperature blue
(TB) = 725 K. lines are subcritical isotherms; the purple isotherm is Tc (305 K):
the Boyle temperature (TB ) = 725 K.
It is clear from Equation (1) that ω ≥ 0, i.e., rigidity must always be positive although it can be zero
in two-phase coexistence regions. Gibbs energy cannot decrease with pressure when T is constant.
From these definitions moreover, not only can there be no “continuity” of gas and liquid but the gas and
liquid states are fundamentally different in their thermodynamic description. Rigidity is determined by
Entropy 2018, 20, 22 4 of 15

It is clear from Equation (1) that ω ≥ 0, i.e., rigidity must always be positive although it can be
zero in two-phase coexistence regions. Gibbs energy cannot decrease with pressure when T is constant.
From these definitions moreover, not only can there be no “continuity” of gas and liquid but the gas and
liquid states are fundamentally different in their thermodynamic description. Rigidity is determined by
number density fluctuations at the molecular level, which have different but complementary statistical
origins in each phase, hence the symmetry [9]. There is a distribution of many small clusters in a gas
with one large void; there is a distribution of unoccupied pockets in the liquid with one large cluster.
The statistical neighborhood properties of unoccupied pockets are the same as occupied sites.
The percolation loci and the critical coexisting densities at Tc for CO2 have been obtained by
the method previously described in detail for Lennard-Jones fluids [5] and argon [8]. Looking at
Figure 2, the coexisting densities of the gas (ρG ) and of liquid (ρL ) and the meso-density gap vary
linearly between the critical temperature (Tc ) and Boyle temperature (TB ). If the mesophase rigidity is
obtainable in terms of TB and TC , we can then represent the density bounds with linear functions and
the rigidity in the mesophase as shown in Figure 3a,b, and incorporate a linear equation of state for the
mesophase that connects with gas and liquid at the percolation bounds PB and PA respectively.

Figure 3. (a) Phase diagram for CO2 : data obtained from NIST Thermophysical Properties compilation
indicating the percolation loci that bound the supercritical liquid and gas states: the green and blue
dashed lines are the gas (PB) and liquid (PA) percolation loci respectively [4–9]; solid red points;
coexisting densities of the gas (ρG ) and of liquid (ρL ) at Tc are indicated by the arrows; (b) rigidity as a
function of temperature in the mesophase.

GAS ρPB (T) = ρG (Tc ) [(TB − T)/(TB − TC )] (5)

LIQUID ρPA (T) = ρL (Tc ) [(TB − T)/(TB − TC )] (6)

Use of original experimental data could result in non-linear Equations (5) and (6) for density
state bounds in a more refined analysis. Given these expressions for the temperature dependence
of the percolation loci, and the rigidity in the mesophase (Figure 3b), in terms of physical constants
belonging to the specific fluid, we can now proceed to represent the thermodynamic surfaces using
virial expansions.
The rigidity in the mesophase varies from kT at TB to zero at Tc . It is linearly dependent on
temperature in the near-critical region up about 1.25 T/Tc but there is an increasing small quadratic
term as seen in Figure 3b. The temperature dependence in the range Tc to TB appears to be logarithmic,
Entropy 2018, 20, 22 5 of 15

but without a scientific explanation, and within the uncertainty of the data, it is best represented by a
simple quadratic
ωT = c1 T* + c2 T*2 (7)

where T* = (T − Tc )/(TB − Tc ) and the constants c1 and c2 as given in Figure 3b.

3. Pressure Surfaces p(ρ,T)


The pressure for any thermodynamic equilibrium state point can be obtained as a function of
density along any isotherm. The gas equation-of-state can be parameterized using the formally exact
Mayer virial expansion [10]. The mesophase pressure increases linearly with density the region 0 > T >
TB and ρPB (T) > ρ < ρPA (T), and for the “liquid” state ρ > ρPA (T), we can use an empirical expansion.
The isothermal equations-of-state for pressure are as follows:

GAS p(ρ,T) = kTρ (1 + b2 (T)ρ + b3 (T)ρ2 + . . . ) (8)

MESO p(ρ,T) = p(ρPB ,T) + ωT (ρ − ρPB ) (9)

LIQUID p(ρ,T) = p(ρPA ,T) + [a1 (ρ − ρPA ) + a2 (ρ − PA )2 + a3 (ρ − ρ PA )3 . . . ] (10)

For the temperature-density region Tc < T < TB and ρ < ρPA (T), Equations (8)–(10) are based
upon the observation that there is a phase transition of the third-order at the mesophase bounds
with a discontinuity in the derivative of ω with respect to ρ, i.e., the third derivative of Gibbs energy
with respect to ρ at constant T along an isotherm. For temperatures above TB , p(ρPA ,T) and ρPA ,
in Equation (10), are zero, whereupon the coefficients an in Equation (10) become the same as the
Mayer virial coefficients bn (T) in Equation (8).
For the supercritical fluid at temperatures above TB , Equation (8) up to order b4 is sufficient to
reproduce the pressure with five-figure accuracy, i.e., within the margin of the original experimental
uncertainty, with a trend-line regression >0.999999, up to temperatures of 1000 K and pressures up
to 50 Mpa. For the gas phase, at all temperatures below TB , the 4th virial coefficient is found to be
essentially zero, within the uncertainty of the experimental data; all higher terms are negligible. Thus,
for the case of CO2 , the equation-of-state of the whole “gas” region of Figure 3 requires just the second
and a third virial coefficients b2 (T) and b3 (T) within the experimental precision.
An overall p(ρ,T) equation of state can be integrated along any isotherm to obtain the Helmoltz
energy function A(T) and hence, via the partition function, all other thermodynamic state functions
and/or derivatives by long-established procedures [11].

3.1. Supercritical CO2


In order to demonstrate a necessity of three separate equations-of-state, we take, as an example
in the first instance, a supercritical isotherm T = 350 K, or T/Tc = 1.15, of CO2 . The critical and Boyle
temperatures for CO2 according to NIST databank [3] are Tc = 305 K and Tc = 725 K. Using the same
method described previously for argon [8], the coexisting densities at Tc are found to be ρc (gas) =
7.771 mol/L and ρc (liq) = 13.46 mol/L. Substituting these physical constant values into equations
(5 to 7) for T = 350 K we obtain the gas and liquid state density bounds ρPB (gas) = 7.305 mol/L and
ρPA (liq) = 12.02 mol/L, and the rigidity ω(350 K) = 0.318 kT (0.925 kJ/mol).
If we try to fit the NIST data for the whole isotherm continuously from 0, just up to only 50 MPa,
using a 6-term polynomial for example, we obtain the result:

p = −0.000008ρ6 + 0.000457ρ5 − 0.009373ρ4 + 0.096245ρ3 − 0.596317ρ2 + 3.459264ρ − 0.189239

with R2 = 0.999125
In this typical continuous overall polynomial fit, the lower order coefficients are scientifically
meaningless. By contrast, by using virial coefficients and known properties of the mesophase bounds,
Entropy 2018, 20, 22 6 of 15

the NIST data for the whole range of three regions can be reproduced with essentially 100% precision
when the gas, liquid and meso-states have different equations-of-state. We assume that there is a
third-order phase transition, i.e., a discontinuity in the third derivative of Gibbs energy at the gas and
liquid-state boundaries in accord with inequalities (2 and 3).
The experimental data from the NIST tabulations is reproduced with the same precision as the
original data can be obtained, i.e., accurate to 5 figures, as evidenced by the mean-squared regression
Entropy 2018, 20, 22 6 of 15
(R2 ) in the trendline polynomial coefficients as given. Also shown in Figure 4 are redefined equations
to
togive
givethe
thevirial
virialcoefficients
coefficientsininEquations
Equations(8)–(10)
(8)–(10)asasshown
shownon onthe
theplots.
plots.The
Thecoefficient
coefficientaa11 in
inthe
the
liquid-state expansion is taken to be equal to the value of ω in the mesophase to effect the
liquid-state expansion is taken to be equal to the value of ωT in the mesophase to effect the third-order
T third-order
discontinuity
discontinuitybetween
betweenthethemesophase
mesophaseand andthe
theliquid
liquidstate.
state.There
Thereisisaasymmetry
symmetrybetween
betweenthe therigidity
rigidityof of
gas
gasandandliquid
liquidon oneither
eitherside
sideofofthe
themesophase
mesophasebut butatatthis
thisstage,
stage,this
thisobservation
observationisisempirical.
empirical.As Asitit
presently
presentlylacks
lacksformal
formaltheory,
theory,weweassume
assumeonly
onlythe
thelinear
linearcoefficient
coefficientininEquation
Equation(10)(10)and
andmake
makeno no
assumptions regarding higher-order coefficients at this
assumptions regarding higher-order coefficients at this stage. stage.

(a) (b) (c)


Figure4.4. Pressure
Pressure equations
equationsp(ρ)
p(ρ)for
forCO
CO2 at a supercritical temperature (350 K or T/Tc = 1.15) as derived
Figure 2 at a supercritical temperature (350 K or T/Tc = 1.15) as
from NIST Thermophysical Properties compilation:
derived from NIST Thermophysical Properties compilation: (a) gas (a)
state;
gas(b) supercritical
state; meso-phase
(b) supercritical and (c)
meso-phase
liquid.
and (c) liquid.

3.2. Supercritical Argon


3.2. Supercritical Argon
Modern research into fundamental equation-of-state science is more often based upon the
Modern research into fundamental equation-of-state science is more often based upon the example
example of argon in the first instance because of the relative simplicity of the spherical two-body
of argon in the first instance because of the relative simplicity of the spherical two-body pair potential
pair potential and the amenability to basic theory. The literature on equation-of-state data to high
and the amenability to basic theory. The literature on equation-of-state data to high temperature and
temperature and pressure is both extensive and accurate to at least five figures except in the vicinity of
pressure is both extensive and accurate to at least five figures except in the vicinity of the critical
the critical temperature (Tc) and pressure (pc) [3,12]. For temperatures above the Boyle temperature
temperature (Tc ) and pressure (pc ) [3,12]. For temperatures above the Boyle temperature there appears
there appears to be just a single continuous fluid state up to the highest pressures measured, which, in
to be just a single continuous fluid state up to the highest pressures measured, which, in the case of
the case of argon, is 1000 MPa, according to NIST [3,12]. If there are no discontinuities along isotherms
argon, is 1000 MPa, according to NIST [3,12]. If there are no discontinuities along isotherms for T > TB ,
for T > TB, the Mayer virial expansion is applicable over the whole accessible density range. The
the Mayer virial expansion is applicable over the whole accessible density range. The expansion for
expansion for the pressure is
the pressure is
p =pkT(ρ + b2+ρb2 2+
= kT(ρ ρ2b+3 ρb33ρ+3 +b4bρ44ρ4. .……)
. ... ) (11)
(11)
andon
and ondifferentiation
differentiationthe
therigidity
rigidityisis
ω = (dp/dρ)T = kT(1 + 2b2ρ + 3b3ρ2 + 4b4ρ3………) (12)
ω = (dp/dρ)T = kT(1 + 2b2 ρ + 3b3 ρ2 + 4b4 ρ3 . . . . . . . . . ) (12)
where bn are the temperature-dependent virial coefficients with conventional definitions. The NIST data
where bn supercritical
for four are the temperature-dependent
isotherms above and virial coefficients
including the with
Boyleconventional
temperaturedefinitions.
(TB) can beThe NIST
precisely
data for four supercritical isotherms above and including the Boyle temperature (T ) can
reproduced using Equation (12) and the virial coefficients determined. From StewartB and Jacobson [13] be precisely
reproduced
we obtain Tusing Equation
B = 414 ± 0.5 for (12) and
argon, the virial
which coefficients
is consistent withdetermined. From
the zero value Stewart
of the secondand Jacobson
virial [13]
coefficient
we obtain T = 414 ± 0.5 for argon, which is consistent with the zero value of the second
obtained here from parameterizations of NIST rigidity values from sound velocity compilations.
B virial coefficient
obtained
Thehere from parameterizations
equation-of-state isothermsofare
NIST rigidity values
presented from5.sound
in Figure Also velocity
given are compilations.
the polynomial
trend-line parameters up to order 4. In this region, the entire experimental range of pressures, from
near zero at ideal gas states at limiting low densities, up to 1000 MPa, can be represented to within
the experimental five-figure precision by a three or at most four-term, closed-virial equation. Using
Equation (12) we can deduce lower-order values of the virial coefficients of argon as listed in Table 1.
The values obtained for b2 and b3 are in agreement with the literature values within the experimental
uncertainties, which can be as high as 20% in the case of b3 [9]. For argon at supercritical temperatures,
Entropy 2018, 20, 22 7 of 15

The equation-of-state isotherms are presented in Figure 5. Also given are the polynomial trend-line
parameters up to order 4. In this region, the entire experimental range of pressures, from near zero at
ideal gas states at limiting low densities, up to 1000 MPa, can be represented to within the experimental
five-figure precision by a three or at most four-term, closed-virial equation. Using Equation (12) we can
deduce lower-order values of the virial coefficients of argon as listed in Table 1. The values obtained
for b2 and b3 are in agreement with the literature values within the experimental uncertainties, which
can be as high as 20% in the case of b3 [9]. For argon at supercritical temperatures, little is presently
known Entropy
experimentally
2018, 20, 22 about b4 and higher coefficients. 7 of 15

Figure 5. Rigidity isotherms, ω(ρ) for argon at temperatures from 700 K down to TB (414 K) as
Figure 5. Rigidity isotherms, ω(ρ) for argon at temperatures from 700 K down to TB (414 K) as obtained
obtained from NIST Thermophysical Properties compilation up to pressures of 500 MPa; all of the
from NIST Thermophysical Properties compilation up to pressures of 500 MPa; all of the EXCEL
EXCEL trendline polynomials of order 4 reproduce the NIST isotherms with total precision, i.e., a
trendline polynomials
regression of R2 =of order up
1.000000 4 reproduce the NIST isotherms with total precision, i.e., a regression of
to 6-figure precision.
R2 = 1.000000 up to 6-figure precision.
Table 1. Argon virial coefficients.

T (K) b2 Table 1. Argon


(10−3 mol/L) b3virial coefficients.
(10−6 mol 2/L2) b4 (10−9 mol3/L3)
700 14.91 0.8191 0.0063
T (K) −3 mol/L) b3 (10−6 mol2 /L 2) b4 (10−9 mol3 /L3 )
600 b2 (10 11.45 0.9029 0.0013
700 500 6.130
14.91 1.0523 0.8191 −0.0082 0.0063
600 414 (~TB) 0.0081
11.45 1.3085 0.9029 −0.0248 0.0013
500 151 (~Tc) −86.02
6.130 2.48 1.0523 0 −0.0082
414 (~TB ) 100 −183.2
0.0081 0 1.3085 −0.0248
151 (~Tc ) −86.02 2.48 0
Lower-order
100 virial coefficients, as defined by Equation (12),
−183.2 0 and calculated from the trend-line
polynomial coefficients in Figure 5, are summarized in Table 1. All the experimental data from the
NIST tabulations are reproduced with maximum precision as evidenced by the mean-squared
Lower-order
regression invirial coefficients,
the trendline as defined
polynomial by (R
coefficients Equation
2 = 1.000000(12), and
to six calculated
decimal from
places in the
every trend-line
case);
polynomial coefficients
the coefficients in Figure
have 5, are summarized
been redefined in Table
to obtain values of the 1. All expansion
virial the experimental data
coefficients frominthe NIST
defined
Equation
tabulations (11). These arewith
are reproduced in line with previous
maximum b2(T) and
precision asbevidenced
3(T) tabulationsby [14]
the within the uncertainties,
mean-squared regression in
which are rather wide, up to a few percent
2 for b 2(T), and up to 25% in the case of b3(T).
the trendline polynomial coefficients (R = 1.000000 to six decimal places in every case); the coefficients
have been redefined to obtain values of the virial expansion coefficients defined in Equation (11). These
4. Critical Isotherms
are in line with previous b2 (T) and b3 (T) tabulations [14] within the uncertainties, which are rather
wide, up toArgon
4.1. a few percent for b2 (T), and up to 25% in the case of b3 (T).
The critical temperature of argon according to NIST is 150.87, the value obtained in a revised
4. Critical Isotherms
analysis of the data of Gilgen et al. is 151.1 ± 0.1. Here we represent the rigidities of the experimental
151 K isotherm from the NIST data tables using standard trend-line polymomials. The coexisting gas
4.1. Argon
and liquid densities as obtained from the original experimental data of Gilgen et al. [14,15] are
Theρc(gas) = 11.89
critical mol/L and ρcof
temperature (liq) = 14.95according
argon mol/L [8]. For
to all
NISTtemperatures
is 150.87,between Tc and
the value the triple point
obtained in a revised
(Ttr), the NIST ω(ρ)T isotherms can be reproduced with 100% precision using just the two-parameter
analysis of the data of Gilgen et al. is 151.1 ± 0.1. Here we represent the rigidities of the experimental
equation. Numerical values of the coefficients b2 and b3 at Tc, obtained from the rigidity isotherms as
illustrated in Figure 6a, are included in Table 1. The trend-line polynomials of order 3 reproduce the
NIST isotherms over the whole range of equilibrium existence with high accuracy.
Entropy 2018, 20, 22 8 of 15

151 K isotherm from the NIST data tables using standard trend-line polymomials. The coexisting gas
and liquid densities as obtained from the original experimental data of Gilgen et al. [14,15] are ρc (gas)
= 11.89 mol/L and ρc (liq) = 14.95 mol/L [8]. For all temperatures between Tc and the triple point
(Ttr ), the NIST ω(ρ)T isotherms can be reproduced with 100% precision using just the two-parameter
equation. Numerical values of the coefficients b2 and b3 at Tc , obtained from the rigidity isotherms as
illustrated in Figure 6a, are included in Table 1. The trend-line polynomials of order 3 reproduce the
NIST isotherms
Entropy 2018,over
20, 22 the whole range of equilibrium existence with high accuracy. 8 of 15
These trend-line equations for ωT (ρ) can be integrated to obtain pressure equations, and rewritten
as: These trend-line equations for ωT(ρ) can be integrated to obtain pressure equations, and rewritten
as:
Gas ρ < ρc (gas) p = kT (ρ + b2 ρ2 + b3 ρ3 ) (13)
Gas ρ < ρc(gas) p = kT (ρ + b2 ρ2 + b3 ρ3) (13)
Meso ρc (gas) < ρ < ρc (liq) p = pc (14)
Meso ρc(gas) < ρ < ρc(liq) p = pc (14)
Liquid ρ > ρc (liq) p = p(ρcl ,T) + kT [a (ρ − ρcl ) + a2 (ρ − ρcl )2 + a3 (ρ − ρcl )3 ] (15)
Liquid ρ > ρc(liq) p = p(ρcl,T) + kT [a (ρ − ρcl) + a2(ρ − ρ ) + a3(ρ − ρ ) ]
cl 2 cl 3 (15)
where ρcl is the critical coexisting liquid density ρc (liq). For argon, the critical pressure in Equation (14)
where
pc = 4.949 MPaρcl is[4].
the Thus,
critical coexisting
we have liquid density ρc(liq). For argon,
an equation-of-state the critical
for argon along pressure in Equation
the whole (14) isotherm,
critical pc
= 4.949 MPa [4]. Thus, we have an equation-of-state for argon along whole of the critical isotherm given
with no divergent singularity. The coefficients in Equations (13) and (15), are obtainable from the
the coefficients in Figure 6a,b for Equation (15), obtainable, for example, using the data in Figure 6b,
rigidity by
data plotted
dividing thein Figure 6a,b,
coefficients nkT.example, by dividing the coefficients an by nkT.
an byfor

(a) (b)
Figure 6. Rigidity isotherms, ω(ρ), for argon along Tc from NIST Thermophysical Properties compilation
Figure 6. Rigidity isotherms, ω(ρ), for argon along Tc from NIST Thermophysical Properties
[3]: (a) gas phase from zero to maximum gas density at critical pressure 4.949 MPa (b) liquid phase
compilation [3]: (a) gas
from minimum phase
liquid densityfrom
from zero to MPa)
pc (4.949 maximum gas The
to 326 MPa. density atalong
rigidity critical pressure
the critical 4.949
divide is MPa
(b) liquid phase
equal from minimum liquid density from pc (4.949 MPa) to 326 MPa. The rigidity along the
to zero.
critical divide is equal to zero.
4.2. Steam and Water
4.2. Steam and
TheWater
evidence for supercritical percolation transitions bounding the existence of water and
steam in the supercritical region, and the parameters describing the mesophase, have all been the
Thesubject
evidence for supercritical
of a previous article withpercolation
an extendedtransitions
debate [16].bounding the existence
For the present purposes, we of water
only needand steam
in the supercritical
values of some region, and
physical the parameters
constants describing
that are given the mesophase,
as an ancillary have[3].
data list by NIST all been the subject of a
previous article with
Figure an extended
7 shows how the debate [16]. in
critical point Forthethe present
T-p plane ispurposes, we only need
thermodynamically definedvalues
by anof some
physicalintersection
constantsof twoare
that percolation
given asloci, using the example
an ancillary data listofbywater.
NISTThe[3].coexisting density curves are
taken from the experimental measurements of the IAPWS International Steam Tables [17]. The
Figure 7 shows how the critical point in the T-p plane is thermodynamically defined by an
supercritical percolation transition points PA and PB intersect at Tc, and continue to define the
intersection of two percolation loci, using the example of water. The coexisting density curves are taken
metastable limit of existence loci, usually referred to as spinodals, of the subcritical gas and liquid
from thewithin
experimental
the two-phasemeasurements
region. Valuesofofthe theIAPWS International
critical physical Steam
constants Tables
for water are: [17]. TheK;
Tc = 647.1 supercritical
pc =
percolation
22.05transition
MPa; criticalpoints
steam PA andρPB
density cG = intersect
13.02 mol/L; at critical
Tc , and continue
water densitytoρcdefine
L = 20.61the metastable
mol/L [3,16]. limit of
Also
existence loci, plotted referred
usually in Figure 7toare
asthe experimentally
spinodals, of the observed spinodals
subcritical [18] that
gas and bound
liquid the regions
within of
the two-phase
region. metastable
Values ofexistence within
the critical the two-phase
physical constantscoexistence
for waterregion at subcritical
are: Tc = 647.1temperatures.
K; pc = 22.05 At T c, the
MPa; critical
percolation loci, PA and PB in Figure 7, cross the critical coexistence line at its extremities to become
steam density ρcG = 13.02 mol/L; critical water density ρcL = 20.61 mol/L [3,16].
the subcritical boundary of the metastable compressed gas (p > psat) and metastable expanded liquid
Also plotted
state in) stability
(p < psat Figure 7limits
are the experimentally
respectively. observed
This behavior spinodals
was also shown for [18] thatargon
liquid bound[8]. the
It isregions
of metastable existence
consistent within the two-phase
with a phenomenological definition coexistence
of PA and PB region at liquid
when both subcritical
and gastemperatures.
have the same At Tc ,
values of the rigidity (ωT) on the same isotherm whereupon (d2p/dρ2)T = 0 at both PA and PB but at
Entropy 2018, 20, 22 9 of 15

the percolation loci, PA and PB in Figure 7, cross the critical coexistence line at its extremities to become
the subcritical boundary of the metastable compressed gas (p > psat ) and metastable expanded liquid
state (p < psat ) stability limits respectively. This behavior was also shown for liquid argon [8]. It is
consistent with a phenomenological definition of PA and PB when both liquid and gas have the same
values of the rigidity
Entropy (ωT ) on the same isotherm whereupon (d2 p/dρ2 )T = 0 at both PA9and
2018, 20, 22 of 15 PB but at
differentEntropy
pressures.
2018, 20,
At these instability points, there is also zero surface tension and consequently
22 9 of 15
no
different pressures. At these instability points, there is also zero surface tension and consequently no
barrier to spontaneous nucleation of steam from water at PA, or water from steam at
barrier to spontaneous nucleation of steam from water at PA, or water from steam at PB, for example.
PB, for example.
different pressures. At these instability points, there is also zero surface tension and consequently no
barrier to spontaneous nucleation of steam from water at PA, or water from steam at PB, for example.

Figure 7. A phase diagram of water and steam in the p-T plane: the red and blue data points and
Figure 7. A phase diagram of water and steam in the p-T plane: the red and blue data points and
dashed lines are the percolation loci state bounds PA (water) and PB (steam) respectively.
dashed lines are the percolation loci state bounds PA (water) and PB (steam) respectively.
The 7.
Figure equations-of-state
A phase diagramfor of steam
water and water
steam along
in the the
p-Tcritical isotherm
plane: the areblue
red and shown
datainpoints
Figureand
8a,b
respectively.
dashed linesInare
both
thecases a quartic
percolation equation
loci is required
state bounds to fit and
PA (water) the data with the
PB (steam) five-figure precision of
respectively.
The equations-of-state for steam
the NIST tables. The fourth and water
virial coefficient along
for steam at the criticalstill
Tc, however, isotherm
appears toare
beshown in in
quite small Figure 8a,b
respectively. In both
itsThe cases
contribution rightaup
quartic
equations-of-state to for equation
the steam
near critical is required
density.
and water We also
along thetonote
fit the data
thatisotherm
critical with
the liquid the
water
are five-figure
critical
shown precision
isotherm
in Figure 8a,b of
extendsThe
the NISTrespectively.
tables. from a critical
In fourth
both casespressure
virial ofequation
22.05–1000
coefficient
a quartic for MPa and
is steam
required can
atto c ,be
Tfit reproduced
however,
the data with with
still six-figure
theappears
five-figure toprecision
be quiteofsmall in
precision
theby the quartic
NIST polynomial.
tables. The fourth virial coefficient for steam at Tc, however, still appears to be quite small in
its contribution right up to the near critical density. We also note that the liquid water critical isotherm
extendsits contribution right up to the near critical density. We also note that the liquid water critical isotherm
from a critical pressure of 22.05–1000 MPa and can be reproduced with six-figure precision by
extends from a critical pressure of 22.05–1000 MPa and can be reproduced with six-figure precision
the quartic polynomial.
by the quartic polynomial.

(a) (b)
Figure 8. Rigidity (ω) for steam and water along the critical isotherm from NIST Thermophysical
Properties compilation [4]: (a) steam from zero to ρcG at Tc (rigidity = 0); (b) liquid phase from minimum
liquid density (ρcL) at pc = 22.05 MPa (ω = 0) to a pressure of 1000 MPa (ω ~ 100 kJ/mol).
(a) (b)
It is clear from the above that the equations-of-state for subcritical isotherms will require a
Figure knowledge
scientific 8. Rigidity (ω) forforms
of the steamfor
and water along thethecritical isotherm fromthe
NIST Thermophysical
Figure 8. Rigidity (ω) for steam and the extension
water alongofthe percolation
critical loci into
isotherm fromsubcritical two-phase
NIST Thermophysical
Properties
region on thecompilation [4]: (a) At
density surface. steam from we
present zerodo ρcG at
to not Tc (rigidity
have = 0); (b) liquid
this information. phase be
It would from minimum
rather unwise
Properties compilation
liquid [4]:
density (ρcL) at pc =(a) steam
22.05 MPa (ωfrom
= 0) zero to ρcG ofat1000
to a pressure Tc MPa
(rigidity = 0);
(ω ~ 100 (b) liquid phase from
kJ/mol).
minimum liquid density (ρcL ) at pc = 22.05 MPa (ω = 0) to a pressure of 1000 MPa (ω ~100 kJ/mol).
It is clear from the above that the equations-of-state for subcritical isotherms will require a
scientific knowledge of the forms for the extension of the percolation loci into the subcritical two-phase
region on the density surface. At present we do not have this information. It would be rather unwise
Entropy 2018, 20, 22 10 of 15

It is clear from the above that the equations-of-state for subcritical isotherms will require a
scientific knowledge of the forms for the extension of the percolation loci into the subcritical two-phase
region on the density surface. At present we do not have this information. It would be rather unwise
to make any assumptions based only upon the known supercritical behavior of the density loci of PA
and PB and rather scanty subcritical spinodals. In the following section, therefore, we will limit the
presentEntropy
extension
2018, 20,of
22 the above to an analysis of the 100 K isotherm of liquid argon, to demonstrate
10 of 15
the applicability of the simple equation-of-state forms Equations (8)–(10) to subcritical gas and liquid
to make any assumptions based only upon the known supercritical behavior of the density loci of PA
state isotherms.
and PB and rather scanty subcritical spinodals. In the following section, therefore, we will limit the
present extension of the above to an analysis of the 100 K isotherm of liquid argon, to demonstrate
5. Subcritical Isotherms
the applicability of the simple equation-of-state forms Equations (8)–(10) to subcritical gas and liquid
state isotherms.
Subcritical Argon
5. Subcritical Isotherms
The 100 K (T/Tc ~2/3) isotherm data for the rigidity from the NIST compilation of gaseous and
Subcritical
liquid argon Argon
are re-plotted in Figure 9. The gas pressure up to the condensation line can be represented
with just theThe
two-term equation with the
100 K (T/Tc ~ 2/3) isotherm datavalue ofrigidity
for the b2 givenfrominthe
Table 1; compilation
NIST b3 is essentially zero.
of gaseous Thus the
and
accurateliquid
equation for re-plotted
argon are argon gas in at lower
Figure temperatures
9. The gas pressure upis simply
to the condensation line can be represented
with just the two-term equation with the value of b2 given in Table 1; b3 is essentially zero. Thus the
accurate equation for argon gas atplower temperatures
gas (100 K) = kTρis(1simply
+ b2 ρ) (16)
pgas (100 K) = kTρ (1 + b2 ρ) (16)
The value obtained from the linear fit of ωT in Figure 8a, is 183.2 × 10−3 mol/L compares
The value obtained from the linear fit of ωT3in Figure 8a, is 183.2 × 10−3 mol/L compares favorably
favorably with an experimental value −187 (cm /mol) obtained in 1958 [13].
with an experimental value −187 (cm3/mol) obtained in 1958 [13]

(a) (b)
Figure 9. Rigidity ω(ρ) for argon along the sub-critical isotherm T = 100 K from NIST [4]: (a) gas
Figure 9. Rigidity ω(ρ) for argon along the sub-critical isotherm T = 100 K from NIST [4]: (a) gas
phase from zero to maximum equilibrium gas coexistence density (0.42 mol/L): (b) liquid phase from
phase from zero to
minimum maximum
liquid equilibrium
equilibrium coexistencegas coexistence
density density
(32.9 mol/L) at the (0.42 mol/L):
coexistence (b) liquid
pressure (MPa) phase
to a from
minimum liquid equilibrium
pressure of 100 MPa. coexistence density (32.9 mol/L) at the coexistence pressure (MPa) to a
pressure of 100 MPa.
The corresponding pressure data for liquid argon at 100 K up to freezing (62 MPa) may not be
continuous in all its derivatives, and hence the equation-of-state may not be as simple as it looks if
The corresponding
there is a higher-orderpressure data forThe
discontinuity. liquid argon
liquid at 100
isotherm K up to
appears tocomprise
freezingtwo (62 linear
MPa)parts
may not be
continuous in allanitsintersection
containing derivatives, at anand hence the
intermediate equation-of-state
density may This
of around 35.0 mol/L. not be as simple
is consistent withasait looks
if there is a higher-order discontinuity. The liquid isotherm appears to comprise two linear parts
previous conjecture [19] that as the crystallization point is approached, the liquid itself may become
another
containing kind of mesophase
an intersection at anas large clusters density
intermediate of orderedofarrangements
around 35.0ofmol/L.
atoms, probably bcc in the with a
This is consistent
case of argon, begin to stabilize. This could be a feature of the equilibrium “liquid” state structure in
previous conjecture [19] that as the crystallization point is approached, the liquid itself may become
the near triple-point density region. One can regard the data in Figure 9b as a fragment of evidence
anotherforkind of mesophase as large clusters of ordered arrangements of atoms, probably bcc in the
this conjecture. We will not consider this intriguing possibility further at this stage. We note,
case of argon,
however, begin
that ifto stabilize.
this is indeedThis could be science,
the underlying a featureanyof the equilibrium
equation-of-state “liquid”
for the state‘pure
equilibrium structure in
the nearliquid’
triple-point density region. One can regard the data in Figure 9b as a fragment
may terminate before the equilibrium crystallization density is reached along subcritical of evidence for
isotherm, below around 115 K in the case of argon.
this conjecture. We will not consider this intriguing possibility further at this stage. We note, however,
that if this is indeed the underlying science, any equation-of-state for the equilibrium ‘pure liquid’ may
Entropy 2018, 20, 22 11 of 15

terminate before the equilibrium crystallization density is reached along subcritical isotherm, below
around 115 K in the case of argon.

6. Energy and Heat Capacities

6.1. Argon Critical Isotherm and Isochores


The only real evidence for a van der Waals-like singularity and a critical point, with continuity
of gas and liquid [1] and universal scaling properties [20], is evidently in historic heat capacity
measurements [21,22]. Here we focus briefly at the equation-of-state for the internal energy U(ρ)T for
the critical isotherm of argon close to Tc (151 K). The NIST data with the trend-line parameters are
shown in Figure 9. The foremost observation is that along this isotherm the gas phase is represented
by a quadratic requiring two coefficients whereas the liquid phase is linear up to a density around
Entropy 2018, 20, 22
35 mol/L. The critical divide is linear by definition in accord with the lever rule or11 proportionality. of 15

The internal energy


6. Energy and forHeatall temperatures below Tc in the two-phase region is defined to be linear
Capacities
by the lever rule since the physical state is of two-phase coexistence. Above the critical divide the three
6.1. Argon Critical Isotherm and Isochores
regions shown in Figure 10, however, look much the same, so we can identify two higher-order phase
The only real evidence for a van der Waals-like singularity and a critical point, with continuity
transitions bounding
of gas and theliquid
mesophase. There
[1] and universal is noproperties
scaling evidence [20],of
is divergent behavior
evidently in historic of slope of U(T)ρ, i.e.,
heat capacity
in the isochoricmeasurements
heat capacity [21,22].
CHere we focus briefly at the equation-of-state for the internal energy U(ρ)T for
v . Isotherms from Tc to 2Tc all show a constant value of (dU/dT)ρ in
the critical isotherm of argon close to Tc (151 K). The NIST data with the trend-line parameters are
the supercritical region,
shown in Figureas 9.seen in Figure
The foremost 11. Along
observation the this
is that along critical
isothermdivide, the isochoric
the gas phase is representedheat capacity is
defined by the lever rule: for T ≤ Tc
by a quadratic requiring two coefficients whereas the liquid phase is linear up to a density around 35
mol/L. The critical divide is linear by definition in accord with the lever rule or proportionality.
The internal energy for all temperatures below Tc in the two-phase region is defined to be linear
Cvthe
by the lever rule since = physical
(dU/dT) stateVis =
of X Cv (liq)
two-phase + (1 − Above
coexistence. X) Cvthe (gas)
critical divide the three (17)
regions shown in Figure 10, however, look much the same, so we can identify two higher-order phase
where X is moletransitions
fraction bounding the mesophase. There is no evidence of divergent behavior of slope of U(T)ρ, i.e.,
of liquid.
in the isochoric heat capacity Cv. Isotherms from Tc to 2Tc all show a constant value of (dU/dT)ρ in
Uhlenbeckthe[20] appears to ashave
supercritical region, seen inbeen
Figurethe firstthe
11. Along tocritical
recognize what
divide, the heheat
isochoric described
capacity is as “surprising
defined by the lever rule: for T ≤ T
Russian results” [21,22] as being supportive of a universal picture of critical phenomena in which the
c

isochoric heat capacity Cv is believed Cv = to


(dU/dT) V = X Cv(liq) + (1 − X) Cv(gas)
diverge logarithmically on both sides of(17) the critical point.
where X is mole fraction of liquid.
When the proceedings of the Washington Conference were published one-year after the event [20],
Uhlenbeck [20] appears to have been the first to recognize what he described as “surprising
the articles by Rowlinson,
Russian results”and by
[21,22] as Fisher, in particular,
being supportive hailed
of a universal picturethe Russian
of critical discovery
phenomena in whichasthea breakthrough
(referenced by Rowlinson though
isochoric heat capacity curiously
Cv is believed to not referenced
diverge logarithmicallybyonFisher),
both sideswith
of thethe implicit
critical point. proclamation
When the proceedings of the Washington Conference were published one-year after the event [20], the
that the van derarticles
Waalsbycritical point, with continuity of liquid and gas, had become established scientific
Rowlinson, and by Fisher, in particular, hailed the Russian discovery as a breakthrough
truth, and universal scaling
(referenced properties
by Rowlinson thoughsimilar
curiouslytonotareferenced
2D-Ising by model lattice
Fisher), with gas would
the implicit describe gas-liquid
proclamation
that the van der Waals critical point, with continuity of liquid and gas, had become established
and indeed all similar “critical point singularities”. In fact, an abundance of experimental evidence
scientific truth, and universal scaling properties similar to a 2D-Ising model lattice gas would describe
against this conjecture
gas-liquid and was already
indeed there
all similar in point
‘critical the singularities’.
literature In [23]
fact,and not supportive
an abundance of experimentalof van der Waals
evidence against this conjecture was already there in the literature [23] and not supportive of van der
hypothesis or universality.
Waals hypothesis or universality.

Figure 10. Internal energy U(ρ) for argon along the isotherm (T = 151 K) using data from NIST [4].
Figure 10. Internal energy U(ρ) for argon along the isotherm (T = 151 K) using data from NIST [4].
Entropy 2018, 20, 22 12 of 15
Entropy 2018, 20, 22 12 of 15

0.6
0.4
0.2

Internal Energy (KJ/mol)


GAS
0
-0.2 2-PHASE
-0.4
-0.6
-0.8 LIQUID
-1
145 146 147 148 149 150 151 152 153 154 155 156
T(K)
(a) (b)
Figure 11. (a) Internal energy isochores, U(T)ρ for argon for a temperature range from near triple
Figure 11. (a) Internal energy isochores, U(T)ρ for argon for a temperature range from near triple point
point to supercritical from NIST [4] showing an expanded window picture of the near critical
to supercritical from NIST [4] showing an expanded window picture of the near critical isochores to
isochores to the right. (b) Magnified data in critical region showing a weak discontinuity in (dU/dT)v,
the right. (b)atMagnified
i.e., Cv, datanoinevidence
Tc but clearly critical region showing
of logarithmic a weak discontinuity
divergence in T(dU/dT)
on either side of c. v , i.e., Cv, at Tc
but clearly no evidence of logarithmic divergence on either side of Tc .
Despite using complex equations-of-state that could accommodate the concept of a critical
Despiteat
density the node
using of a parabola,
complex the NIST thermodynamic
equations-of-state data bank is the
that could accommodate still concept
valid as “experimental”
of a critical density
evidence that the 1965 Washington proclamation of ‘a universality theory’ was fundamentally incorrect.
at the node of a parabola, the NIST thermodynamic data bank is still valid as “experimental” evidence
It was based upon a misinterpretation of the isochoric heat capacity Cv for T < Tc, and spurious results
that the 1965 Washington proclamation of ‘a universality theory’ was fundamentally incorrect. It was
at, or millikelvins greater than, Tc [24,25]. Along any isochore within the critical divide there can be no
baseddivergence
upon a misinterpretation
of the isochoric heat of capacity.
the isochoric
For T heat
> Tc acapacity
divergence Cvoffor CvTwould
< Tc , and
be a spurious results
contradiction of at,
or millikelvins greater than, IfTcone
laws of thermodynamics. [24,25].
could Along anyadd
reversibly isochore
heat (Qwithin the critical divide there can be no
rev) to a classical Gibbs fluid system that
divergence
does noofwork
the isochoric heat capacity.
with no increase For T > Tc aitdivergence
in the temperature, would be imply, of Cvforwould be a that
example, contradiction
Qrev i.e., of
laws enthalpy,
of thermodynamics. If one could reversibly add heat
or (Qrev/T) i.e., entropy, are not state functions [26]. (Q rev ) to a classical Gibbs fluid system
that does no work with no increase in the temperature, it would be imply, for example, that Qrev i.e.,
6.2. SF
enthalpy, or6 near
(QrevCritical Isochors
/T) i.e., entropy, are not state functions [26].
Sulphur hexafluoride has a special importance in this scientific debate; its critical temperature
6.2. SF 6 near
and Critical
pressure areIsochors
amenable to experiments at near ambient conditions. It has been the subject of a
microgravity experiment aboard the NASA Space Shuttle [27] that appeared to confirm the erroneous
Sulphur hexafluoride has a special importance in this scientific debate; its critical temperature
result obtained some 25 years earlier by Voronel and his colleagues [24,25]. We have obtained the
and pressure are amenable to experiments at near ambient conditions. It has been the subject of a
original published data points of the space shuttle experiments (see acknowledgements) and have
microgravity experiment aboard the NASA Space Shuttle [27] that appeared to confirm the erroneous
plotted them alongside the U(T)V equation-of-state from the NIST data bank parameterized as a simple
resultquadratic
obtainedforsomeT < T25 years earlier by Voronel and his colleagues [24,25]. We have obtained the
c and linear for T > Tc equations in Figure 12a. The equations for U(T) at the same
original
density as the space shuttle μgofexperiment
published data points the spacecan shuttle experiments
be differentiated (seethe
to give acknowledgements)
isochoric heat capacities andCvhave
plotted
in them
Figurealongside the U(T)
12b. In contrast to V equation-of-state
both the historic Russianfromresults
the NISTfor data
argon,bank
and parameterized
the space shuttle as“C
a simple
v”

valuesfor
quadratic also
T< plotted
Tc and in linear
Figure for12b,Tthere
> Tc isequations
no divergence of the 12a.
in Figure isochoric
The heat capacity
equations eitheratbelow
forCvU(T) the same
or above
density as theTspace
c, as suggested,
shuttle µg forexperiment
example, in alternative near critical equations
can be differentiated to give the of isochoric
state with aheat
crossover
capacities
to the divergent form in the vicinity of T c [28].
Cv in Figure 12b. In contrast to both the historic Russian results for argon, and the space shuttle “Cv ”
We can see from Figure 12 that the isochoric heat capacities reported in the Russian experiments,
values also plotted in Figure 12b, there is no divergence of the isochoric heat capacity Cv either below
and the space shuttle experiments, as relied upon by all the principle proponents of universality for
or above Tc , as suggested, for example, in alternative near critical equations of state with a crossover
several decades [20], are a misinterpretation of the correct thermodynamic definition of the isochoric
to theheat
divergent form in the vicinity of Tc [28].
capacity Cv below the critical temperature. If the equations-of-state for U(T) in Figure 12a are
We
essentially correct,Figure
can see from 12 shuttle
the space that theresult,
isochoric
as wellheat capacities
as the reported
argon results for “Cinv”the
mustRussian experiments,
be wrong in the
and the space
region shuttle
above experiments,
Tc. One explanationas relied
may upon
be that theby all the principle
experimental proponents
measurements of universality
on argon [21] were for
several decades
carried out on[20], are a misinterpretation
continuously of theHeat
stirred steady-states. correct thermodynamic
capacity data obtaineddefinition
for stirredof the isochoric
samples in
closed adiabatic calorimeters cannot be construed as equilibrium thermodynamic
heat capacity Cv below the critical temperature. If the equations-of-state for U(T) in Figure 12a are data for the
determination
essentially correct, of thefluid
spaceequations-of-state
shuttle result, as in the
wellclose proximity
as the of Tc, either
argon results for “C below or above
v ” must Tc, forin the
be wrong
any isochore along the critical divide. In the case of the SF 6 microgravity experiments [22], if there
region above Tc . One explanation may be that the experimental measurements on argon [21] were
carried out on continuously stirred steady-states. Heat capacity data obtained for stirred samples
in closed adiabatic calorimeters cannot be construed as equilibrium thermodynamic data for the
determination of fluid equations-of-state in the close proximity of Tc , either below or above Tc , for
Entropy 2018, 20, 22 13 of 15

Entropy 2018,
any isochore 20, 22 the critical divide. In the case of the SF microgravity experiments [22],
along 13 ofif
15 there
6
was no stirring, there must be some other reason for the discrepancy seen in Figure 12, that is
was no stirring, there must be some other reason for the discrepancy seen in Figure 12, that is presently
presently inexplicable.
inexplicable.

(a) (b)
Figure 12. (a) Internal energy isochores, U(T)ρ for SF6 for a temperature range from near triple point
Figure 12. (a) Internal energy isochores, U(T)ρ for SF6 for a temperature range from near triple point
to supercritical from NIST Thermophysical Properties compilation [4]; (b) Isochoric heat capacities
to supercritical from NIST Thermophysical Properties compilation [4]; (b) Isochoric heat capacities
calculated from the equations in Figure 1a, i.e., (dU/dT)ρ alongside the original space shuttle heat
calculated from
capacity datathe equations in Figure 1a, i.e., (dU/dT)ρ alongside the original space shuttle heat
points.
capacity data points.
7. Conclusions
7. Conclusions
From the foregoing analysis, we conclude that the pressure equation-of-state, and hence all
other
Fromthermodynamic state functions,
the foregoing analysis, can be expressed
we conclude simply inequation-of-state,
that the pressure terms of a few physical constants
and hence all other
belonging to the fluid, and coefficients in virial expansions. The van der Waals equation-of-state
thermodynamic state functions, can be expressed simply in terms of a few physical constants belonging
survived for about 50 years until, increasingly, there was evidence that it could not account for the
to the fluid, and coefficients in virial expansions. The van der Waals equation-of-state survived for
experimental data unless it was extended to include more and more terms and adjustable parameters.
aboutModern
50 years until, increasingly, there was evidence that it could not account for the experimental
equations-of-state with a large number of terms and adjustable parameters can reproduce
data experimental
unless it wasp-V-Textended
data totohighinclude
precisionmoreforand moreengineering
chemical terms and applications,
adjustable parameters.
but they mayModern
no
equations-of-state
longer reflect an with a large number
underlying physical of termsinand
science the adjustable
spirit of vanparameters
der Waals. can reproduce experimental
p-V-T data Wetohave
highfound
precision forequation-of-state
that the chemical engineeringfor the lowapplications, but they
density gas phases, in may no longer
all three reflect an
temperature
regions physical
underlying T > TB, TB science
>T > Tc, and
in theT <spirit
Tc, canofallvanbe represented
der Waals. by the first two or three terms of a Mayer
virial
We haveexpansion.
found ForthatT the
> TBequation-of-state
a single virial equation can low
for the represent the whole
density isotherms
gas phases, upthree
in all to thetemperature
highest
recorded
regions T > TBpressures,
, TB >T >which
Tc, and in Tthe
<Tcase, of argon
can all be isrepresented
100 MPa. For by supercritical
the first twotemperatures
or three below
terms of atheMayer
c
Boyle temperature the gas phase can be represented by a term virial expansion to within the accuracy
virial expansion. For T > TB a single virial equation can represent the whole isotherms up to the highest
that the original pressures reported by NIST [3]. The virial equations-of-state are fundamental to all
recorded pressures, which in the case of argon is 100 MPa. For supercritical temperatures below the
atomic and molecular fluids and appear to be accurate up to second order, i.e., to quadratic terms for
Boylealltemperature
temperaturesthe gas phase
below TB. Wecan willbe represented
report by a term virial
detailed comparisons and expansion to within
virial coefficients the case
for the accuracy
that the
CO2, and also some preliminary comparisons with experimental data and NIST equations, for the to all
original pressures reported by NIST [3]. The virial equations-of-state are fundamental
atomic and molecular
exemplary fluids
fluid argon, andcourse.
in due appear to be accurate up to second order, i.e., to quadratic terms for
all temperatures below
At this stage TBdevelopment
in the . We will report detailed alternative
of science-based comparisons formsand virial
of the coefficientswe
equation-of-state forhave
the case
CO2 ,not
and attempted
also some to parameterise
preliminarythe second and third
comparisons withvirial coefficients as
experimental datafunctions
and NISTof temperature.
equations,We for the
note, however,
exemplary therein
fluid argon, is due
evidence that b2(T) may be non-analytic at the Boyle temperature. It can be
course.
represented
At this stage byin
twothedifferent series expansions,
development at least in
of science-based the case of aforms
alternative simple ofLennard–Jones model, we
the equation-of-state
above and below TB [29,30]. There are also some indications that b3(T) exhibits a maximum at or near
have not attempted to parameterise the second and third virial coefficients as functions of temperature.
Tc. Also, we find that for simple fluids, e.g., argon, below TB, the coefficient b4(T) appears to become
We note, however, there is evidence that b2 (T) may be non-analytic at the Boyle temperature. It can be
vanishingly small. A more accurate determination of lower-order virial coefficients from original
represented by two data
thermodynamic different
oughtseries
now toexpansions,
become a researchat leastpriority.
in the case of a simple Lennard–Jones model,
above and below TB [29,30]. There are also some indications that b3 (T) exhibits a maximum at or near
Tc . Also, we find that We
Acknowledgments: for wish
simple fluids,
to thank e.g., Sadus
Richard argon,of below TB ,University
Swinburne the coefficient b4 (T) appears
for providing toheat
the original become
capacity data
vanishingly points
small. Afrom
moretheaccurate
space shuttle experiments [27]of
determination re-plotted in Figure
lower-order 12b. coefficients from original
virial
thermodynamic data ought now to become a research priority.
Entropy 2018, 20, 22 14 of 15

Acknowledgments: We wish to thank Richard Sadus of Swinburne University for providing the original heat
capacity data points from the space shuttle experiments [27] re-plotted in Figure 12b.
Conflicts of Interest: The author declares no conflict of interest.

References
1. Van der Waals, J.D. “Over de Continuiteit van den Gas-en Vloeistoftoestand” (On the Continuity of the Gas
and Liquid State). Ph.D. Thesis, University of Leiden, Leiden, The Netherlands, 1873.
2. Span, R.; Wagner, W. A New Equation of State for Carbon Dioxide Covering the Fluid Region for temperatures
from the Triple Point to 1100 K at Pressures up to 800 MPa. J. Phys. Chem. Ref. Data 1996, 25, 1509–1594.
[CrossRef]
3. NIST Thermophysical Properties of Fluid Systems. 2017. Available online: http://webbook.nist.gov/
chemistry/fluid/ (accessed on 20 May 2017).
4. Woodcock, L.V. Observations of a thermodynamic liquid-gas critical coexistence line and supercritical phase
bounds from percolation loci. Fluid Phase Equilib. 2012, 351, 25–33. [CrossRef]
5. Heyes, D.M.; Woodcock, L.V. Critical and supercritical properties of Lennard-Jones Fluids. Fluid Phase Equilib.
2013, 356, 301–308. [CrossRef]
6. Magnier, H.J.; Curtis, R.C.; Woodcock, L.V. Nature of the supercritical Mesophase. Nat. Sci. 2014, 6, 797–807.
[CrossRef]
7. Woodcock, L.V. Percolation Transitions and Fluid State Boundaries. CMST 2017, 23, 281–294. [CrossRef]
8. Woodcock, L.V. Gibbs density surface of fluid argon: Revised critical parameters. Int. J. Thermophys. 2014, 35,
1770–1777. [CrossRef]
9. Woodcock, L.V. Thermodynamics of Gas-Liquid Criticality: Rigidity Symmetry on Gibbs Density Surface.
Int. J. Thermophys. 2016, 37, 24. [CrossRef]
10. Mayer, J.E.; Mayer, M.G. Statistical Mechanics; Wiley: New York, NY, USA, 1940.
11. Michels, A.; Geldermans, M.; de Groot, S.R. Determination of thermodynamic properties from laboratory
compressibility data. Physica 1946, 12, 105. [CrossRef]
12. Tegeler, C.; Span, R.; Wagner, W. A new equation of state for argon covering the fluid region for temperatures
from the melting line to 700 K at Pressures up to 1000 MPa. J. Phys. Chem. Ref. Data 1999, 28, 779–850.
[CrossRef]
13. Stewart, R.B.; Jacobsen, R.T. Thermodynamic properties of argon from triple point to 1200 K with pressures
up to 1000 MPa. J. Phys. Chem. Ref. Data 1989, 18, 639–678. [CrossRef]
14. Gilgen, R.; Kleinrahm, R.; Wagner, W. Measurement and correlation of the (pressure, density, temperature)
relation of argon: I. The homogeneous gas and liquid regions in the temperature range from 90 to 300 K at
pressures up to 12MPa. J. Chem. Thermodyn. 1994, 26, 383–398. [CrossRef]
15. Gilgen, R.; Kleinrahm, R.; Wagner, W. Measurement and correlation of the (pressure, density, temperature)
relation of argon II Saturated-liquid and saturated vapour densities and vapour pressures along the entire
coexistence curve. J. Chem. Thermodyn. 1994, 26, 399–413. [CrossRef]
16. Woodcock, L.V. Gibbs density surface of water and steam: 2nd debate on the absence of van der Waals’
critical point. Nat. Sci. 2014, 6, 411–432. [CrossRef]
17. Wagner, W.; Kretzschmar, H.-J. IAPWS-IF’97 International Steam Tables: Properties of Water and Steam; Springer:
Berlin, Germany, 2008.
18. Shamsundar, N.; Lienhard, J.H. Equations of state and spinodal lines: A review. Nucl. Eng. Des. 1993, 141,
269–287. [CrossRef]
19. Finney, J.L.; Woodcock, L.V. Renaissance of Bernal’s random close packing and hypercritical line in the
theory of liquids. J. Phys. 2014, 26, 463102. [CrossRef] [PubMed]
20. Green, M.S.; Sengers, J.V. Critical Phenomena. In Proceedings of the Conference on Phenomena in the
Neighborhood of Critical Points, Washington, DC, USA, 5–8 April 1965; National Bureau of Standards:
Washington, DC, USA, 1965.
21. Bagatskiı̌, M.I.; Voronel, A.V.; Gusak, B.G. Measurements of isochoric heat capacities of near-critical argon.
Sov. Phys. JETP 1963, 16, 517.
22. Voronel, A.V.; Chorkin, Y.R.; Popov, V.A.; Simkin, V.G. Measurements of isochoric heat capacities of
near-critical argon. Sov. Phys. JETP 1964, 18, 568.
Entropy 2018, 20, 22 15 of 15

23. Rice, O.K. Critical phenomena. In Thermodynamics and Physics of Matter; Rossini, F.D., Ed.; Oxford University
Press: London, UK, 1955; pp. 419–500.
24. Sengers, J.V.; Anisimov, M.A. Comment on “Gibbs density surface of fluid argon: Revised critical properties”
by L. V. Woodcock. Int. J. Thermophys. 2015, 36, 3001. [CrossRef]
25. Woodcock, L.V. On the interpretation of near critical heat capacity measurements. Int. J. Thermophys. 2017,
38, 139. [CrossRef]
26. Sandler, S.I.; Woodcock, L.V. Historical observations on the laws of thermodynamics. J. Chem. Eng. Data
2010, 55, 4485. [CrossRef]
27. Haupt, A.; Straub, J. Evaluation of the isochoric heat capacity measurements at the critical Isochore of SF6
performed during the German Spacelab Mission D-2. Phys. Rev. E 1999, 59, 1975. [CrossRef]
28. Kostrowicka-Wyczalkowska, A.; Sengers, J.V. Thermodynamic properties of Sulphur hexafluoride in the
critical region. J. Chem. Phys. 1999, 111, 1551. [CrossRef]
29. Eu, B.C. Exact analytic second virial coefficient for the Lennard-Jones Fluid. arXiv 2009, arXiv:0909.3326.
30. Heyes, D.M.; Pieprzyk, S.; Rickayzen, G.; Branka, A.C. The second virial coefficient and critical point
behaviour of the Mie potential. J. Chem. Phys. 2016, 145, 084505. [CrossRef] [PubMed]

© 2018 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like