You are on page 1of 8

Journal of Molecular Liquids 295 (2019) 111625

Contents lists available at ScienceDirect

Journal of Molecular Liquids


journal homepage: www.elsevier.com/locate/molliq

R-crossing method applied to fluids interacting through variable


range potentials
Jaime Jaramillo-Gutiérrez, J. Luis López* , José Torres-Arenas
División de Ciencias e Ingenierías Campus León, Universidad de Guanajuato, A.P. E-143, C.P. 37150 León, Guanajuato, México

A R T I C L E I N F O A B S T R A C T

Article history: The R-crossing method of thermodynamic geometry is applied to reproduce the coexistence curves of fluid
Received 21 May 2019 systems described by hard-core Yukawa and hard-core Mye-Type interactions whose range can be var-
Received in revised form 1 August 2019 ied. Connection between the range of the potential and the validity of the method is studied. Even when
Accepted 24 August 2019 scaling relations suggest a dependence on the range, we found explicitly the quantitative dependence for
Available online 2 September 2019
two varying range potentials. Using the saturation pressures, it is possible to assign a percentage P(k) of
measuring how far we can go below the critical point when reproducing the coexistence curve. It is found
Keywords: that there is a close relation between the range of the potential and P(k). Such relation assures that the
Thermodynamic geometry
larger the range of the potential, the deeper we can go below the critical point. Our results, together with
Fluctuation theory
the known low |R| limit, represent two independent criteria to restrict the applicability of the R-crossing
R-crossing method.
method. Namely, at what extent we can accurately reproduce the coexistence curve when we know the
range of the intermolecular potential of the thermodynamic system involved.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction present. It has been shown that this is correct for a large number of
thermodynamic systems [3]. Other properties of R have been found;
Several geometric approaches have been constructed in order to i) the values of |R| represent the sizes of organized structures at
assign geometric properties to thermodynamic spaces mainly in the mesoscopic scales, information which is not provided by the macro-
framework of Riemannian Geometry, in which the distance between scopic thermodynamics or the microscopic statistical mechanics,
points in the space of thermodynamic states can be measured with ii) the sign of R is related to the nature of the interaction, attractive or
the use of the invariant distance squared ds2 = gab dxa dxb and where repulsive, iii) the curvature scalar diverges at the point where a phase
the metric coefficients gab can be consistently defined [1-4]. Those transition occurs, iv) the equality of the values of R in the coexist-
geometrical approaches to thermodynamics have been successful ing phases allows to reproduce the coexistence curve. The procedure
in deriving thermodynamic properties from geometrical quantities, related to this last property is known as the R-crossing method [6-8]
where the curvature scalar related to these spaces plays an important and it uses the equation of state (EOS) of the system analyzed on one
role [3,5-7]. hand, and the curvature scalar of thermodynamic geometry R on the
We will refer mainly to the Ruppeiner’s approach where the other. The main idea is to represent the curvature scalar as a function
starting point was the fluctuation theory of statistical mechanics of volume R(T, V) or pressure R(T, P), this can be done with the help
[2,3]. Using this particular point of view, fluctuation theory allows to of the equation of state setting aside the gas and liquid branches of
establish a clear relation between the curvature scalar and the corre- the curvature scalar. The points of intersection of these two branches
lation length defining the volume where the fluctuating theory loses at constant temperature, in terms of pressure, represent the satu-
3
its validity, namely |R| ∝ n . According to this particular approach, ration pressures that are equal for the two phases. The set of these
it has been conjectured that whenever the curvature scalar is differ- points for isotherms near the critical temperature reproduce, to a
ent from zero, an interaction between the particles of the system is good extent, the coexistence curve. This powerful method allows to
calculate the correlation volume in terms of thermodynamic prop-
erties and calculate coexistence curves accurately in the asymptotic
critical region [6].
* Corresponding author.
E-mail addresses: jaramillogj2011@licifug.ugto.mx (J. Jaramillo-Gutiérrez),
The hypothesis behind the application of the R-crossing method
lopezjl@ugto.mx, jl_lopez@fisica.ugto.mx (J.L. López), jtorres@fisica.ugto.mx was first established in Ref. [7] and regarding fluid systems, it has
(J. Torres-Arenas). been applied for instance for the Lennard-Jones fluid in Refs. [8,9]

https://doi.org/10.1016/j.molliq.2019.111625
0167-7322/© 2019 Elsevier B.V. All rights reserved.
2 J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625

where it was tested the dependence of the curvature scalar R on 2. Thermodynamic geometry overview
the intermolecular strength by using the Weeks-Chandler-Anderson
(WCA) ansatz. The method has also been applied for magnetic The geometrical model proposed by G. Ruppeiner was origi-
systems and an idealized liquid system in Ref. [10]. In these pre- nally inspired in thermodynamic fluctuation theory [2,3]. In this
vious works, we find that the R-crossing method reproduces to a framework, the metric is related with the fluctuation probability p
good extent the coexistence curve near the critical point and the that depends on the quadratic invariant (l)2 where the following
robustness of these results provide a way to predict first order first proportionality follows from the entropy in terms of probability
transitions.  
In this work, we will apply the R-crossing method to varying V
p ∝ exp − (l)2 , (2)
range potentials. The coexistence curve reproduced by the R-crossing 2
method will be compared with the curve that arises from following
the standard thermodynamic procedure, that makes use of the val- where
ues of pressure p(v) and chemical potential l(v) for volumes in the
liquid vl and gas vg phases respectively. The coexistence curve can be (l)2 = g11 (x1 ) + 2g12 x1 x2 + g22 (x2 )2 . (3)
realized by solving the equations
The xa (a = 1, 2, . . . n) represent the extensive parameters of
entropy, and xa = xa − xa0 is the distance of each parameter with
p(vg ) − p(vl ) = 0 , (1)
respect to the reference parameter state xa0 . The form of metric ele-
l(vg ) − l(vl ) = 0 . ments gab depend on the representation, namely the corresponding
thermodynamic potential. Mainly two potentials are used; entropy
S(U, N, V) in terms of internal energy, particle number and volume on
We will apply the R-crossing method, namely reproduce the
one side and free energy A(T, N, V) = U − TS in terms of tempera-
coexistence curve, to fluids interacting through potentials of vari-
ture, particle number and volume on the other. For the case of the
able range; particularly for hard-core (HC)-Yukawa and HC-Mie-type
free energy which is the potential that we will use in this work, the
potentials. Our intention is to see at what extent the method could
metric elements are [2,6]
depend on the range of the potential. The coexistence curve can
be reproduced accurately near the critical point, but it is not clear
1 ∂ 2A 1 ∂ 2A
how far below the critical point we can go in the sense that we can g11 = − 2
, g22 = − , g12 = g21 = 0. (4)
kTV ∂ T kTV ∂ N2
still trust the data coming from the R-crossing method. It has been
observed that the points of the coexistence curve coming from the Here, we have fixed the volume hence the free energy will depend
method begin to separate from the standard curve at a certain value on two parameters only. We have already mentioned in the intro-
of temperature [6-8]. The purpose of our work is then two folding, duction that the R-crossing method proposed in Ref. [7] allows to
first we want to establish a criterion from which we can determine reproduce the coexistence curve for a region near the critical point,
a certain percentage of how far we can go below from the critical mainly because of the proportionality between the curvature scalar
temperature with the coexistence curve still being accurate and sec- R and the correlation length and the commensurate R theorem [6]. In
ond, we want to determine the dependence of this percentage on the order to explain and exemplify how the method works, in the next
range of the interacting potential. section, we will apply the R-crossing method to reproduce the coex-
Although the R-crossing method is powerful, it has been shown istence curve of a Van der Waals fluid. We will compare the curve
that when the values of |R| are comparable to the molecular vol- that follows from the R-crossing method and the curve that is con-
ume vg , the equality between the values of R in the vapor and liquid structed using the standard thermodynamic method. Afterwards, we
phases begin to separate. Along with this, the coexistence curve in will apply the same method to potentials of variable range in order
this region loses accuracy [6,7]. This is known as the low |R| limit to establish a relation between the parameter defining the range
which is explicitly represented by the criterion, |R|/vg  1. We will and the results of applying the R-crossing method to reproduce the
show that the accuracy of the coexistence curve reproduced with coexistence curve.
the R-crossing points compared with the points calculated with the
traditional method also depend on the range of the potential con- 3. R-crossing method
sidered. Moreover, we are able to define a criterion to quantify the
disagreement between the two curves. For the purpose of clarifying the methodology that we will follow
This work is structured as follows, first in Section 2, we briefly in the application of the R-crossing method, first we review in great
review the thermodynamic geometry tools needed to compute the detail the application of the method to the Van der Waals fluid [7]
curvature scalar and in Section 3, we will review in great detail, all to further proceed with other systems interacting with potentials of
the steps involved in the R-crossing method in order to reproduce variable range like the Yukawa and Mie-type ones.
the coexistence curve of the corresponding fluid. We will further
apply the R-crossing method to particular fluids interacting through 3.1. R-crossing method for Van der Waals system
potentials of variable range, namely HC-Yukawa and HC-Mie-type
potentials 1 . As we are interested in the role of the range, we will test The starting point is finding the free energy per volume f = A/V
the method for three different ranges for each type of potential, three which in this case is given by [14]
for the Yukawa and three for the Mie ones. The EOS is obtained with  
the statistical associating fluid of hard core with attractive potentials q
f = qkT ln + qke(T) − qkT ln(1 − bq) − akq2 , (5)
and variable range (SAFT- VR) [11-13]. In Section 4, we will show our q0
results and Sec. 5 will be devoted to conclusions.  
where e(T) = cv T ln TT + 40 and a, b, q0 are constants. The metric
0
components are correspondingly

1 qcv 1 2a
From now on, we will refer to these potentials just as Yukawa and Mie. The hard- gTT = , gqq = − . (6)
core nature of the potentials will be explicit in our procedure T2 q(1 − bq)2 T
J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625 3

Notice from Eq. (4) that the Boltzmann constant k cancels (we did
not take k = 1). In subsequent equations, we use k to reduce the
free energy as in Eq. (13) below. The reduced curvature scalar is then
given by

(3V ∗ − 1) [F(V ∗ , T ∗ ) + G(V ∗ , T ∗ )]


R∗ (T ∗ , V ∗ ) = , (7)
6cv H(V ∗ , T ∗ )

where

F = 4cv T ∗ V ∗2 (1 − 9cv V ∗ + 27V ∗2 − 27V ∗3 + 8T ∗ V ∗3 ), (8)


G = 9V ∗ (1 − 9V ∗ + 27V ∗2 − 27V ∗3 + 24T ∗ V ∗4 )
H = (4T ∗ V ∗3 − 9V ∗2 − 1)2 .

We have used the critical values to define the following reduced


quantities, P∗ = P/Pc , T∗ = T/Tc , V∗ = V/Vc where

a 8a
Pc = , Tc = , Vc = 3Nb, (9)
27b2 27kb

and the curvature scalar is reduced with the constant b, namely


R∗ = R/b. We need to make a distinction between the curvature
scalar in the liquid (Rl ) and gas (Rg ) phases respectively. In order to
do that, we will use the EOS that clearly defines the intervals of vol-
ume where the fluid is in these two phases which we will call Vl and
Vg . The reduced pressure by the equation of state is given by

8T ∗ 3
P∗ = − 2. (10)
3V ∗ − 1 V∗

For constant temperature, two solutions of this cubic equation


are associated with the volumes in the liquid Vl∗ (P ∗ ) and gas Vg∗ (P ∗ )
phases respectively. Notice that these volumes can be expressed
explicitly in terms of pressure, hence we can use Vl∗ (P ∗ ) and Vg∗ (P ∗ )
to write the curvature scalars associated with the liquid and gas
phases in terms of pressure, namely we can express Eq. (7) in terms
of pressure and temperature as R(Vl∗ , T ∗ ) → Rl (P ∗ , T ∗ ) and R(Vg∗ , T ∗ ) →
Rg (P ∗ , T ∗ ). We are interested in the points where the plot of both Fig. 1. (a) Isotherms of the curvature scalar R∗ for the Van der Waals fluid. The temper-
scalars Rl (P∗ , T ∗ ) and Rg (P∗ , T ∗ ) intersect each other (see Fig. 1). The ature has been reduced with the critical temperature. This curve shows the tendency
of |R| to diverge as it approaches the critical temperature. (b) Plot of one point of
point in pressure P∗ where the intersection is found represents a sat-
intersection of the curvature scalars R∗l and R∗g for T ∗ = 0.98. The set of all points of
uration pressure and in these points of intersection it is considered intersection at different temperatures reproduce the coexistence curve.
that both phases coexist, since R is proportional to the correlation
length for temperatures near the critical temperature Tc . For the
problem in hand, R can be expressed analytically and the points of
intersection can be found solving the following equation
1
Traditional Method
Rl (T ∗ , P ∗ ) − Rg (T ∗ , P ∗ ) = 0. (11) R-crossing

In Fig. 2, we show in a single plot, the standard coexistence curve 0.8


and the one followed by the application of the R-crossing method.
We can observe in Fig. 2 that there is a temperature at which the
curve predicted by the R-crossing method begin to deviate from the
T*

0.6
standard coexistence curve. That is related to the fact that the pro-
portionality between the curvature scalar and the correlation length
is valid in the neighborhood of the critical point. This point of sep-
aration can also be observed from the saturation pressures as it is 0.4

showed in Fig. 3.

3.2. R-crossing for HC-Yukawa interactions 0.2

0 0.5 1 1.5 2 2.5 3


In this subsection, we will apply the R-crossing method to repro- 1/V*
duce the coexistence curves of systems with potentials with a well
defined range parameter, namely the Yukawa potential. We are basi- Fig. 2. Superposed coexistence curves for the Van der Waals fluid, the one followed
cally interested in understanding the role that the range of the poten- by the standard method (dashed line) and the one followed by the R-crossing method
tial plays, in reproducing the coexistence curve. Let us then consider (circles).
4 J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625

1 where g represents the packing fraction g = qps 3 /6, and

1 − geff /2
0.9 gHS (1; geff ) = . (17)
(1 − geff )3

0.8 The form of the function defined in Eq. (17) will be used below
with a different argument. The parametrization for geff is given by
T*

0.7
geff (g, k) = c1 g + c2 g2 . (18)

0.6
The fluctuation term a∗1 has the form

0.5 a1 (k)∗ = aVdW


1 gHS (1; g∗ ), aVdW∗
1 = −6g4k−1 , (19)

0.4 where the parametrization of g∗ is in this case given by g∗ (g, k) =


0 0.2 0.4 0.6 0.8 1
P* d1 g + d2 g2 and d1 and d2 together with the coefficients of the
parametrization in Eq. (18) are given in Appendix 1. Finally, the form
Fig. 3. Plot of the two saturation pressures for the Van der Waals fluid, the standard
of a2 for the free energy in Eq. (14) is
one (squares) and the ones coming from the R-crossing method (circles).
1 HS ∂ a∗1 (k)
aY2 = 4K g . (20)
2 ∂g
a hard sphere fluid whose interaction is given by the following
potential The hard sphere isothermal compressibility in terms of g is


⎨∞ if r < s, (1 − g)4
K HS = . (21)
0(r) = −k sr −1
( ) (12) 1 + 4g + 4g2
⎩− 4e
r if r > s .
( )
s
Now that we have specified aY1 and aY2 , the free energy and the
We will follow the same procedure as the one described in the SAFT-VR EOS can be found with the use of Eqs. (14) and (13) and con-
previous section. For the system in consideration, we will use a par- sequently we are now able to compute the relevant thermodynamic
ticular form for the free energy that was proposed in Ref. [11], where quantities from the free energy
a version of the statistical associating fluid theory for chain molecules
of hard core with attractive potentials and variable range (SAFT- 64
f (g, T ∗ ) = gT ∗ aY , (22)
VR) were developed. The different contributions to the free energy ps 3
were evaluated according to the Wertheim perturbation theory [15].
A general form of the SAFT reduced Helmholtz free energy aY (Y where the temperature T ∗ is reduced with its critical value
stands for the Yukawa potential in this case) takes into account sev- T ∗ = T/Tc . Notice that the last expression for free energy is naturally
eral contributions [11], but we will consider this expression up to the written in terms of the packing fraction g instead of density q as it
monomer contribution which is given by was the case for the Van der Waals system, therefore we will further
calculate R(g, T ∗ ) in terms of these variables for convenience. Using
Eq. (22) we have that the (reduced) pressure and chemical potential
A AId AM
aY = = aId + aM = + , (13) are
NkT NkT NkT
ps 3 ∂ aY
where both terms correspond to the ideal gas and monomer con- P ∗ (g, T ∗ ) = P(g, T ∗ ) = g2 T ∗ , (23)
tributions respectively. Following the standard theory of hard-core
64 ∂g
systems [16,17] and the high-temperature expansion of Ref. [18], the
and
monomer free energy can be expressed as a series in the inverse
temperature b = 1/kT as
1 ∂ aY
l ∗ (g, T ∗ ) = l(g, T ∗ ) = T ∗ aY + g . (24)
M HS 2
4 ∂g
a =a + ba1 + b a2 . (14)
Following the same guiding lines as in the case of the Van der
a1 and a2 are the first two perturbation terms associated with the Waals system, at this point, we need to calculate the curvature scalar
attractive energy −40 [17], and a2 will be function of the so called R(g, T ∗ ) and this can be done directly by the use of Eq. (22) but we
fluctuation term a∗1 that will be defined below. An extensive study for need to solve the following issue. Notice that in this case we cannot
these contributions for different potentials is given in Ref. [11]. For write an analytical expression for the inverse function g(P∗ ) but, it is
the case of the Yukawa interactions and for ranges between 1.1 ≤ possible to do this inversion numerically, thus, we can use the pres-
k ≤ 4, we have explicitly sures in the liquid and vapor phases to find Rl (P∗ , T ∗ ) and Rg (P∗ , T ∗ ).
Temperature, pressure and packing fraction have been reduced with
aY1 = aVdW
1 gHS (1; geff ) (15) their corresponding critical values (see the tables in the Results
section.) For the case of the Yukawa and Mie systems, we will fur-
with ther reduce the curvature scalar with s 3 , namely R∗ = R/s 3 . In
the same way, points in pressure where both scalars coincide for the
same temperature are considered as saturation pressures where both
aVdW = −12g4(k−1 + k−2 ), (16)
1 phases coexist and consequently, for temperatures near the critical
J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625 5

point, the coexistence curve can be reproduced with the R-crossing in Eq. (27), aHS is the reference hard sphere free energy [23] with
method. In the section of Results, we will analyze the coexistence packing fraction g = qps 3 /6
curves followed by the usual thermodynamic method and those
with the R-crossing method. A crucial observation related with the
4g − 3g2
range of the potentials involved will be noticed immediately, but this aHS = . (28)
(1 − g)2
fact will be pointed out in the aforementioned section. Let us first
apply the R-crossing method to the family of Mie-Type interactions.
Related to these interactions are well defined potentials in terms also The two first perturbation terms, and the ones we need to find the
of a parameter defining their range; parameter that can be varied to free energy, are given by
test again the R-crossing method for these potentials.
 ∞
a1 = 2pq gsHS (r)u1 (r)r2 dr, (29)
s
3.3. R-crossing for Mie-type potentials  ∞
a2 = −pqK HS (1 + w) gsHS (r)(u1 (r))2 r2 dr,
s
In this section, we mainly explain how to obtain the EOS in order
to further apply the R-crossing method to systems with Mie type
where we have for the compressibility in this case
interactions. These potentials represent a generalized interaction
pair potential which includes at once, a Maxwellian repulsive contri-
bution u(r) ∝ r−a and a Newtonian attractive contribution u(r) ∝ r−b (1 − g)4
K HS = . (30)
where the parameters a and b characterize the variable range of 1 + 4g + 4g2 − 4g3 + g4
the repulsive and attractive interactions respectively [12,13,19-21].
These kind of potentials are also known as generalized Lennard-Jones The detailed approximation method applied to calculate a1 and
potentials and such name is clearly understood from the analytical a2 can be found in Ref. [13]. The final form of a1 is given explicitly by
form of Mie-Type potentials which is the following

a1 =C xk0 [aS1 (g; k) + B(g; k)] (31)



∞ if r < s, − C xa0 [aS1 (g; a) + B(g; a)].
0(r) =  a  k  (25)
C 4 sr − sr if r ≥ s .
The calculation of these terms involve the first order terms
aS1 (g; k) of the free energy of a system of hard-core Sutherland parti-
where C is a constant given by
cles of range k. Those contributions are evaluated within the SAFT-VR
and the use of the mean value theorem MVT [13]. We have for
   k instance for the first order term
a a a−k
C= (26)
a−k k  
1 1 − geff (g; k)/2
aS1 (g; k) = −124g , (32)
k−3 (1 − geff (g; k))3
In this potential, the range is given basically by the parameter
k. In the following analysis we will fix a = 12 and we will take
and the parametrization for geff (g; k) is given by
the three following values for k varying the range of the potential;
k = 5, k = 6, k = 8. As it has been stressed in the previous
sections, we need the EOS for the corresponding thermodynamic sys- geff (g, k) = c1 (k)g + c2 (k)g2 + c3 (k)g3 + c4 (k)g4 , (33)
tem, consequently we need an expression for the free energy in order
to calculate thermodynamic properties and the curvature scalar. For
where the ci (i = 1, 2, 3, 4) in the last expression are also showed
the particular case of the Lennard-Jones fluid, the free energy and
in Appendix 1. Once these contributions are given, it is possible
EOS was treated in Ref. [22], and a highly accurate EOS correspond-
to complete the free energy required to compute the components
ing to the generalization of the SAFT-VR for the Mie potentials has
of the metric. The free energy expression we need is given by Eq.
been presented in Refs.[12,13]. We will use the expressions derived
(22) with the reduced Yukawa free energy aY replaced by the cor-
in this reference to apply the R-crossing method to this fluid. The free
responding aMie = AMie /NkT. Analogously, the expressions for the
energy is A(N, V, T) = Aid (N, V, T) + Ar (N, V, T) where residual free
pressure and chemical potential are given by Eqs. (23) and (24). In
energy can be written as the following series
order to calculate the metric components it is possible to use the
relation between density and packing fraction to write the reduced
free energy (Eq. (22)) in terms of density and by the use of the EOS
ares = aHS + ba1 + b2 a2 + · · · (27)
we have now we are able to find the corresponding Rl (q∗ , T ∗ ) and
Rg (q∗ , T ∗ ). The density has been reduced with s 3 as q∗ = qs 3 .
Different approximations to the real Mie potential are mainly Finally, the same numerical inversion procedure is followed to find
used in order to construct the free energy of the respective fluid [12, Rl (P∗ , T ∗ ) and Rg (P∗ , T ∗ ) and the coexistence curve can be reproduced
13], one of those considers an effective system of hard spheres of by intersecting their isotherms.
effective diameter d depending on the temperature. This approxi-
mated potentials are usually called soft-core Mie. We will consider 4. Results
instead the corresponding approximation for hard-core Mie poten-
tials in which the diameter of the spheres stays fixed, namely s does Now we proceed to analyze our results comparing the coexis-
not depend on temperature. Nevertheless, the perturbation terms in tence curves followed by the standard procedure and those obtained
Eq. (27) can be computed with the expressions given in Refs.[12,13] with the R-crossing method for the potentials of variable range.
with constant s, for instance the radial distribution functions (RDF) In the following tables appear the values of ranges tested for the
gsHS (r) needed to calculate the ai factors in Eq. (27). In the first term Yukawa and Mie potentials and the corresponding critical values of
6 J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625

these thermodynamic systems, calculated by the numerical resolu- From the saturation pressures, we provide a criterion to define
tion of the corresponding theoretical EOS. when the method ceases to be accurate; this is when the points in
the saturation pressure coming from the R-crossing method deviate
more than 5% from the points of the standard saturation pressure
Critical values coming from the EOS of the standard theory curve. Based on this criterion, It is possible to assign a percentage
P(k), depending on the range parameter, of how far we can go from
Yukawa Tc gc Pc P(k)
the critical point in reproducing the coexistence curve.
k = 1. 8 1.2711 0.1827 0.1422 15 % Starting with the coexistence curves of the systems described
k = 3 0.8054 0.1837 0.1115 12.8 %
by the Yukawa potentials and for the ranges given, we have that
k = 4 0.6380 0.2015 0.0999 10.5 %
we can descend below the critical temperature (on the coexis-
tence curves) the following percentages, keeping the points between
the saturation pressures coming from the R-crossing and the tra-
Critical values coming from the EOS of the standard theory
ditional method differing by less than 5%; P(k = 4) = 10.5%,
P(k = 3) = 12.8% and P(k = 1.8) = 15% respectively. Up to
Mie Tc gc Pc P(k) now, we observe the tendency of increasing P(k) with decreasing k,
k = 5 1.6273 0.2412 0.0751 16.5 % or in other words, the larger the range of the potential the deeper
k = 6 1.2492 0.2492 0.0596 13 % we can reproduce the coexistence curve with the R-crossing method
k = 8 0.9632 0.3135 0.0550 8.5 % of Thermodynamic Geometry. In order to prove (or disprove) the
previous observation, we have applied the method to the Mie-Type
interactions which are also well defined in terms of their variable
range. In this case, we have tested the range parameters; k = 8,
In Figs. 4 and 5, we show the coexistence curves derived from
k = 6 and k = 5 and the same behavior is observed. The cor-
the R-crossing method for Yukawa and Mie systems respectively,
responding percentages for the tested Mie potentials are given by
superposed with the ones coming from the standard procedure. We
P(k = 8) = 8.5%, P(k = 6) = 13% and P(k = 5) = 16.5%. This
observe that the curve obtained by the R-crossing method repro-
duces the coexistence curve accurately near the critical point, but at behavior clearly shows that there exist a dependence of the repro-
a certain temperature it begins its deviation from the standard curve. duction of coexistence curves of the R-crossing method on the range
This deviation was already observed in Ref. [8]. It also can be noticed of the potential.
that the temperature where the deviation begins is closer to the crit-
ical temperature as the range of the potential decreases. In Fig 6, 5. Conclusions
we show also the saturation pressures, those related to the Yukawa
potentials on the left column and the ones corresponding to the Mie In this work, we have applied the R-crossing method of Rupeiner’s
potentials on the right column. In these plots we also observe a devi- thermodynamic geometry to systems described by potentials of
ation of the points coming from the R-crossing method and those variable range; specifically for HC-Yukawa and HC-Mie-Type poten-
coming from the standard thermodynamic procedure. It is also easier tials. This method allows to reproduce the coexistence curve of the
to observe in these figures the temperature at which the deviation thermodynamic system using the curvature scalar R of thermody-
begin to become significative. namic geometry. For that purpose, the free energy of these systems is

Yukawa
1
0.8
0.6
0.4
0 1 2 3 4
1
0.8
T/Tc

0.6
0.4
0.2
0 1 2 3 4
1
0.8
0.6
0.4
0.2
0 1 2 3
η/ηc

Fig. 4. Coexistence curves plotted together, the ones constructed through the R-crossing method (circles) and the standard ones (dashed lines) for the HC Yukawa interactions
for the following ranges; k = 1.8 (top), k = 3 (middle) k = 4 (bottom).
J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625 7

Mie

0.9

0.8

0 0.5 1 1.5 2 2.5 3


1
T/Tc

0.9

0.8

0.7
0 0.5 1 1.5 2 2.5 3
1
0.9
0.8
0.7
0.6
0 0.5 1 1.5 2 2.5
ρ/ρc

Fig. 5. Coexistence curves plotted together, the ones constructed through the R-crossing method (circles) and the standard ones (dashed lines) for the HC Mie-Type interactions
for the following ranges; k = 5 (top), k = 6 (middle), k = 8 (bottom).

needed in order to compute the metric components. For both poten- curvature scalar is range dependent. We made the analysis for three
tials, Yukawa and Mie, the EOS was obtained with the statistical different ranges for the Yukawa and Mie potentials. For all these
associating fluid theory of hard core with attractive potentials and systems, we found that the larger the range of the potential is, the
variable range (SAFT- VR) [11-13]. farther we can go below the critical temperature in reproducing,
The purpose of our analysis considering variable range poten- to a certain accuracy, the coexistence curve. Based in our analy-
tials was to establish if the accuracy of this description through the sis, it is possible to establish a well defined criterion of how far in

Yukawa Mie
1 1
0.9 0.9
0.8
0.8
0.7
0.7
0.6
0.5 0.6

0.4 0.5
0.3 0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1
0.9 0.9
0.8 0.8
T/Tc

0.7
0.7
0.6
0.6
0.5
0.4 0.5
0.3 0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1
0.9 0.9
0.8
0.8
0.7
0.7
0.6
0.5 0.6

0.4 0.5
0.3 0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
P/Pc P/Pc

Fig. 6. Saturation pressures corresponding to the systems with hard-core Yukawa interactions on the left side and to those with hard-core Mie-Type interactions on the right.
From top to bottom; k = 1.8 (top), k = 3 (middle) k = 4 (bottom) for HC-Yukawa potentials and k = 5 (top), k = 6 (middle), k = 8 (down) for HC Mie-Type potentials. The
curves with circles correspond to the saturation pressures given by the R-crossing method and the dashed lines to the standard thermodynamic saturation pressures.
8 J. Jaramillo-Gutiérrez, J. López and J. Torres-Arenas / Journal of Molecular Liquids 295 (2019) 111625

temperature we can go from the critical point in order to trust the The four ci
s in Eq. (33) are given by
thermodynamic information given by the thermodynamic geome-
⎛ ⎞ ⎛ ⎞⎛ ⎞
try analysis which will clearly depend on the range of the potentials c1 e11 e12 e13 e14 1
involved. ⎜ c2 ⎟ ⎜ e24 ⎟ ⎜ −1 ⎟
⎜ ⎟ = ⎜ e21 e22 e23 ⎟⎜ k ⎟ (A.5)
In Refs. [6,7], it was observed that when the values of |R| are ⎝ c3 ⎠ ⎝ e31 e32 e33 e34 ⎠ ⎝ k−2 ⎠
comparable to the molecular volume vg , the agreement between c4 e41 e42 e43 e44 k−3
the R
s in the vapor and liquid phases begins to fail and conse-
quently the R-crossing method, which is based on the commensurate and in this case
R theorem, begins to be inaccurate. This is known as the low |R| limit.
The values of R that we report based on the thermodynamics fol- e11 = 0.81096, e12 = 1.7888, (A.6)
lowed by the EOS we consider, are far from this limit. Consequently,
e13 = −37.578, e14 = 92.284,
our results represent an independent criterion to restrict the valid-
ity of the R-crossing method when a range dependent potential is e21 = 1.0205, e22 = −19.341,
considered. Moreover, based on our methodology it is possible to e23 = 151.26, e24 = −463.50,
quantify the disagreement between the coexistence curves which e31 = −1.9057, e32 = 22.845,
is necessary if one chooses to calculate them with the R-crossing
method. e33 = −228.14, e34 = 973.92,
e41 = 1.0885, e42 = −6.1962,
e43 = 106.98, e44 = −677.64.
Acknowledgments

J. L. López acknowledge partial support by CONACyT, UG and References


PRODEP grant 511-6/18-8876. J. Torres-Arenas was partially sup-
ported by Convocatoria Institucional de Investigación Científica grant [1] F. Weinhold, Metric geometry of equilibrium thermodynamics, J. Chem. Phys.
007/2019. 63 (65) (1975) 2484–2488. 559(1976).
[2] G. Ruppeiner, Thermodynamics: a Riemannian geometric model, Phys. Rev. A
20 (1979) 1608.
[3] G. Ruppeiner, Riemannian geometry in thermodynamic fluctuation theory, Rev.
Appendix A. Values in the effective packing fraction Mod. Phys. 67 (1995) 605.
parametrizations [4] H. Quevedo, Geometrothermodynamics, J. Math. Phys. 48 (2007) 013506.
[5] H. Quevedo, F. Nettel, C.S. Lopez-Monsalvo, A. Bravetti, Representation invari-
ant geometrothermodynamics: applications to ordinary thermodynamic sys-
The constant c1 and c2 in the parametrization Eq. (18) are given
tems, J. Geom. Phys. 81 (2014) 1.
by [6] G. Ruppeiner, Thermodynamic curvature from the critical point to the triple
point, Phys. Rev. E 86 (2012) 021130.
[7] G. Ruppeiner, A. Sahay, T. Sarkar, G. Sengupta, Thermodynamic geometry,
⎛ ⎞ ⎛ ⎞⎛ ⎞ phase transitions and the Widom line, Phys. Rev. E 86 (2012) 052103.
c1 e11 e12 e13 1
⎝ ⎠=⎝ ⎠ ⎝ k−1 ⎠ , [8] H.-O. May, P. Mausbach, G. Ruppeiner, Thermodynamic curvature for attractive
(A.1) and repulsive intermolecular forces, Phys. Rev. E 88 (2013) 032123.
c2 e21 e22 e23 k−2 [9] H.-O. May, P. Mausbach, Riemannian geometry study of vapor-liquid phase
equilibria and supercritical behavior of the Lennard-Jones fluid, Phys. Rev. E 85
(2012) 031201.
[10] A. Dey, P. Roy, T. Sarkar, Information geometry, phase transitions, and the
where the matrix values are
Widom line: magnetic and liquid systems, Physica A 392 (2013) 6341.
[11] A. Gil-Villegas, A. Galindo, P.J. Whitehead, S.J. Mills, G. Jackson, Statistical asso-
ciating fluid theory for chain molecules with attractive potentials of variable
e11 = 0.900678, e12 = −1.50051, e13 = 0.776577, (A.2) range, J. Chem. Phys. 106 (1997) 4168.
[12] T. Lafitte, D. Bessieres, M.M. Piñeiro, J.-L. Daridon, Simultaneous estimation
e21 = −0.314300, e22 = 0.257101, e23 = −0.0431566. of phase behavior and second-derivative properties using the statistical asso-
ciating fluid theory with variable range approach, J. Chem. Phys. 124 (2006)
024509.
The parameters for d1 , d2 for the parametrization of g∗ (g, k) in Eq. [13] T. Lafitte, A. Apostolakou, C. Avendaño, A. Galindo, Accurate statistical associat-
(19) are ing fluid theory for chain molecules formed from Mie segments, J. Chem. Phys.
139 (2013) 154504.
⎛ ⎞ [14] L.D. Landau, E.M. Lifshitz, Statistical Mechanics, Butterworth-Heinemann,
1 1980.
⎛ ⎞ ⎛ ⎞
d1 e11 e12 e13 e14 e15 ⎜ k−1 ⎟ [15] M.S. Wertheim, Thermodynamic perturbation theory of polymerization, J.
⎜ ⎟ Chem. Phys. 87 (1987) 7323.
⎝ ⎠=⎝ ⎠ ⎜ k−2 ⎟ (A.3)
⎜ ⎟ [16] J.A. Barker, D. Henderson, Perturbation theory and equation of state for fluids
d2 e21 e22 e23 e24 e25 ⎝ k−3 ⎠ II. A successful theory of liquids, J. Chem. Phys. 47 (1967) 4714.
k−4 [17] J.A. Barker, D. Henderson, What is “liquid”? Understanding the states of matter,
Rev. Mod. Phys. 48 (1975) 587.
[18] R.W. Zwanzig, High-temperature equation of state by a perturbation method. I.
with Nonpolar gases, J. Chem. Phys. 22 (1954) 1420.
[19] G. Mie, Zur kinetischen Theorie der einatomigen Krper, Ann. Phys. 316 (1903)
657.
e11 = 0.989601, e12 = −0.872203, e13 = 0.320808, (A.4) [20] E.A. Grneisen, Z. Elektrochem, Angew., Phys. Chem. 17 (1911) 737.
[21] E.A. Grneisen, Z. Elektrochem, Ann. Phys. 344 (1912) 257.
e14 = e15 = 0. [22] J.K. Johnson, J.A. Zollweg, K.E. Gubbins, The Lennard-Jones equation of state
e21 = −0.0119152, e22 = −1.24029, e23 = 2.41636 revisited, Mol. Phys. 78 (3) (1993) 591.
[23] N.F. Carnahan, K.E. Starling, Equation of state for nonattracting rigid spheres, J.
e24 = −2.01922, e25 = 0.647565. Chem. Phys. 51 (1969) 635.

You might also like