Sie sind auf Seite 1von 9

The evolution of multicomponent systems at high pressures

Part II.¤ The Alder–Wainwright, high-density, gas–solid phase transition of the


hard-sphere Ñuid

J. F. Kenney

Joint Institute of the Physics of the Earth, Russian Academy of Sciences ; Gas Resources
Corporation, 15333 JFK Boulevard, suite 160, Houston, T X 77032, USA.
E-mail : JFK=alum.MIT .edu

Received 3rd March 1999, Accepted 24th May 1999

The thermodynamic stability of the hard-sphere gas has been examined, using the formalism of scaled particle
theory (SPT), and by applying explicitly the conditions of stability required by both the second and third laws
of thermodynamics. The temperature and volume limits to the validity of SPT have also been examined. It is
demonstrated that SPT predicts absolute limits to the stability of the Ñuid phase of the hard-sphere system, at
all temperatures within its range of validity. Because SPT describes Ñuids equally well as dilute gases or dense
liquids, the limits set upon the system stability by SPT must represent limits for the existence of the Ñuid phase
and transition to the solid. The reduced density at the stability limits determined by SPT is shown to agree
exactly with those of that estimated for the AlderÈWainwright, high-density gasÈsolid phase transition in a
hard-sphere system, at a speciÐc temperature, and closely over a range of more than 1000 K. The temperature
dependence of the gasÈsolid phase stability limits has been examined over the range 0.01 KÈ10 000 K. It is
further shown that SPT describes correctly the variation of the entropy of a hard-core Ñuid at low
temperatures, requiring its entropy to vanish as T ] 0 by undergoing a gasÈsolid phase transition at Ðnite
temperature and all pressures.

1. Introduction system is examined by applying all the conditions of stability


required by the three laws of thermodynamics.
The high-pressure, supercritical, gasÈliquid phase transition The purpose of this article is to demonstrate the ability of
was Ðrst reported speciÐcally for the real Ñuid normal octane, SPT to predict the AlderÈWainwright, high-density, gasÈsolid
n-C H , in a previous article which will be referred to hence- phase transition in a hard-sphere gas. There has been
8 18
forth as I.1 That particular molecular species had been chosen common misunderstanding that SPT only reproduces the
for analysis, perhaps fortuitously, for reasons of an availability pressureÈdensity relationship of a hard-sphere system from
of data. In I, it was asserted that the high-pressure, Ðrst-order, low densities until, and through, the AlderÈWainwright phase
gasÈliquid phase transition is a geometric property of Ñuids transition, but fails to predict or reproduce that transition.
which must exist independently of the details of any attractive Such misunderstanding is corrected in the following sections
component of the intermolecular potential. This argument where is explicitly described the prediction of the AlderÈ
was supported in I by showing that the gasÈliquid phase tran- Wainwright transition by SPT when constrained by general
sition would occur in a system composed of imaginary stability conditions. The density at which the ÑuidÈsolid phase
““ isomers ÏÏ of n-octane which, while maintaining all other transition occurs in a hard-sphere system is shown to depend
properties of that Ñuid, might possess a characteristic tem- upon its temperature. Furthermore, the temperature-
perature, as deÐned by the formalism of the simpliÐed per- dependence of the ÑuidÈsolid phase transition predicted by
turbed hard chain theory (SPHCT), of one millikelvin, T * \ SPT requires that the Ñuid undergoes such transition at all
0.001 K. Such a characteristic temperature would describe pressures as the temperature decreases to very low tem-
a Ñuid e†ectively without any long-range, attractive van der peratures, precisely in keeping with general thermodynamic
Waals component to its intermolecular potential, and which considerations. Additionally, the latent change of entropy of
would be a gas at all temperatures and low pressures. That the ÑuidÈsolid transition predicted by SPT for hard sphere
““ isomer ÏÏ of n-octane was shown to undergo the gasÈliquid systems at low temperatures is shown to be consistent with
phase transition at almost the same pressure as the real Ñuid. experimental values of the melting entropy for volatile spher-
In order to demonstrate rigorously that the high-pressure, ical molecules.
supercritical, Ðrst-order, gasÈliquid phase transition is a geo- Because SPT uses a factored partition function which is
metric property of Ñuids, in this article and a following one, obtained from application of the BakerÈCampbellÈHausdor†
the stability of hard-core Ñuids is explicitly investigated. expansion, it is a high-temperature formalism. Therefore, the
Because the formalism of scaled particle theory (SPT) was low-temperature limits of the validity of SPT have been exam-
applied extensively in I as the reference system for the joint ined explicitly and are described below.
partition function, the competence of that formalism to The stabilities of the one-dimensional, ““ hard-rod ÏÏ, and
describe the stability of a hard-core system has been investi- two-dimensional, ““ hard-disk ÏÏ systems have also been investi-
gated thoroughly. In both this and the article to follow, the gated and are described in two Appendices. When the general
ability of SPT to predict phase transitions in a hard-core stability conditions, as used with the three-dimensional, hard-
sphere Ñuid, are applied to the one-dimensional, ““ hard-rod ÏÏ
¤ Part I : ref. 1. system, scaled particle theory predicts no instability against a

Phys. Chem. Chem. Phys., 1999, 1, 3277È3285 3277


ÑuidÈsolid phase transition of the AlderÈWainwright type. with any single state of a given system. As such, these condi-
When those same stability conditions are applied to the two- tions are local, a property evidenced by their expression as
dimensional, ““ hard-disk ÏÏ system, it is shown to manifest a di†erential inequalities.
ÑuidÈsolid, ““ freezing ÏÏ transition of the AlderÈWainwright The second law constitutes a global constraint upon the sta-
type at reduced densities somewhat greater than 0.8. bility of a system, as evidenced in its statement by De
DonderÏs inequality :
2. Background
A(p, T , Mn N) \ [ ; lÕ kÕ [ 0 # Mi, o, / ½ (R 0 P)N. (2)
The hard-core Ñuid has been extensively studied, especially j i, p i, p
i, p, Õ
since the discovery of the high-density, gasÈsolid phase tran- In eqn. (2), the summations are taken over all components, i,
sition by Alder and Wainwright in the hard-sphere system.2 A phases, /, and chemical reactions, o, which involve possible
phase transition in a hard-sphere system had been suggested initial (reagent, R) and Ðnal (product, P) states of the system ;
earlier in theoretical work by Kirkwood, who argued that an the mathematical entity, A(p, T , Mn N), is the thermodynamic
ordered conÐguration should appear in the system at a j
affinity, so deÐned by De Donder.8 The non-local condition
reduced density of approximately 0.67.3 Later, Born and set forth by eqn. (2) places the constraint upon any possibility
Green developed the same conclusions.4 However, the investi- that a system in one state might evolve spontaneously into
gations by Kirkwood (and also those by Born and Green) another such that its internal entropy must increase. While De
involved extrapolations of an assumed two-particle corre- DonderÏs inequality sets a necessary condition for stability, it
lation function for dense Ñuids from its form in the thermody- sets the sufficient condition for phase separation only when
namic regime of dilute gases. The calculations of Kirkwood the states involved refer unequivocally to separated phases.
(and Born) could not predict accurately even the fourth virial The third law also sets additionally a stringent condition for
coefficient for simple Ñuids ; and such failure cast severe the physical existence of any state : The total entropy of the
doubts about the validity of their predictions of a phase tran- system must be positive.9h12
sition. The explicit solution by Alder and Wainwright of the
equations of motion of assemblages of hard spheres in 1957, S P 0. (3)
and their demonstration of the phase transition, removed all The inequality (3) is, in fact, only a weaker corollary of the
uncertainties about its existence.2 third law, and does not address the behavior of the entropy as
During the same period while Alder and Wainwright were temperature approaches absolute zero. Strictly, inequality (3)
conducting their computer experiments, Reiss et al.5h7 were may be considered a consequence of BoltzmannÏs classical
developing the formalism of SPT for the analysis of hard-body deÐnition, S \ k ln(W ), in which W represents the number of
systems which applied a radically di†erent perspective to sta- B
conÐgurations which the system can assume, consistent with
tistical mechanics and employed a mathematical expression of its thermodynamic constraints. Because W is a natural
the arcane branch of mathematics designated statistical number, P1, the inequality (3) results. The third law states
geometry.5 The description of the thermodynamic properties that entropy is an absolute entity, with no arbitrary additive
of the hard-sphere gas by SPT represents one of the few constants, and always greater than zero. Unlike De DonderÏs
exactly-solvable problems in quantum statistical mechanics. A inequality, a failure of NernstÏs theorem, inequality (3), for any
particularly important aspect of SPT, which obtains from its phase of a system constitutes both a necessary and sufficient
property as an exact, geometric description, is that it generates condition for a phase transition.
an absolute solution. There are no adjustable parameters or
additive constants to the solution to the problem of the hard-
core gas given by SPT. This property of SPT is used directly
4. Scaled particle theory (SPT)
below to predict, unequivocally, the AlderÈWainwright high- SPT was Ðrst developed by Reiss et al.,5h7,13h18 from argu-
density, gasÈsolid phase transition, both at high temperatures ments of statistical geometry. Carnahan and Starling19 later
and as the temperature of the system approaches absolute derived the SPT equation of state, augmented by the addi-
zero. tional, negative term involving the cubic power of the reduced
density, [g3/(1 [ g)3, by summing explicitly the series devel-
3. The thermodynamic conditions for stability oped from the computed virial coefficients of the hard-sphere
Each of the three laws of thermodynamics states di†erent con- gas. SPT (with the CarnahanÈStarling term) generates the
ditions for the stability of a system. All must be satisÐed for explicit canonical partition function :
the stability of a system.
The Ðrst and second laws together develop the conditions )(V , T ; N)SPT \ )IG)hvc \
1 V NAB
for stability of a single state of a single-component, single- N ! j3
phase system which require that its isothermal compressibility
be greater than zero, that its isometric speciÐc heat be greater
C A
] exp [
BD
g(4 [ 3g) N
. (4)
(1 [ g)2
than zero, that its isobaric speciÐc heat be greater than its
isometric speciÐc heat, that the Ðrst derivative of its chemical In eqn. (1), the ““ thermal de Broglie wavelength,ÏÏ j \
potential with respect to pressure be positive and the second h/J(2pmk T ), where h is PlanckÏs constant and k Boltz-
derivative be negative : B B
mannÏs constant ; T is the absolute temperature and m the

[
A B A BA B
1 dV
[ 0;
dp dT
[ 0;
n
t
molecular mass ; g is the ““ reduced-density ÏÏ parameter,
g \ V */V \ n(p/6)p3/V \ o(p/6)p3, in which p is the diameter
V dp dT dV
A BA B T V p of the hard-spheres, n the molar abundance.
t
dp dT t The factored partition function, eqn. (4), constitutes implicit
[0 t application of the BakerÈCampbellÈHausdor† lemma,20h24
dT dV
A B V
A B p t
dS dS o (1) Q \ trSexp([bH(p, q))T \ trSexp([b(T (p) ] X(q)))T
T \ C [ 0; T \C [C t
dT V dT p Vt \ trSexp([bT (p))TtrSexp([bX(q))T
A B
dk A BV
[ 0;
d2k
\0
p
t
t
] tr
TA1[
b2
[T , X] ] É É É
BU
dp dp2 t 2
T,n T,n p
The conditions for stability set out in eqn. (1) are concerned D trSexp([bT (p))TtrSexp([bX(q))T. (5)

3278 Phys. Chem. Chem. Phys., 1999, 1, 3277È3285


(In eqn. (5), the Hamiltonian, H, has been expressed conven-
tionally in terms of its respective kinetic and potential energy
operators, T (MpN) and )(MqN) of its system momenta MpN, and
spatial MqN coordinates ; and b \ 1/k T .) Neglect of the third
B
factor on the right side of eqn. (5) constitutes a high-
temperature approximation. Thus, from the form of its parti-
tion function, SPT is a formalism valid only at high
temperatures ; and its low temperature limits are investigated
explicitly in Section 6.
The SPT canonical partition function generates the Helm-
holtz free energy :
C A B
FSPT \ Nk T [ ln
V
]1]
g(4 [ 3g) D
\ FIG ] Fhc ;
B Nj3 (1 [ g)2
(6)
of which the isothermal volume derivative gives the system
pressure :

p \ Nk T
1 C
]
1 2g(2 [ g) D
\ pIG ] phc. (7)
B V V (1 [ g)3

Fig. 1 Entropy of a hard-sphere gas at 300 K as a function of the


5. The absolute entropy of a hard-sphere system, reduced-density parameter, g.
and the Alder–Wainwright, high-density, gas–solid
phase transition
The hard-core contribution to the partition function as
described by SPT is purely entropic. This property results gAW \ 0.67114.25 The position of the gasÈsolid phase tran-
intrinsically from the arguments from geometric statistics sition estimated by the Ðxed value of the reduced-density
which underlie SPT and are indi†erent to temperature. There- parameter, gAW, is seen to lie very closely, within a few
fore, the only contribution to the internal energy of a hard- percent, to that determined by the vanishing of the SPT
sphere Ñuid is obtained from the ideal gas factor in its entropy. In fact and as may be observed in Fig. 1, the estimate
partition function, such that USPT \ 3/2Nk T . When the of the AlderÈWainwright transition using the Ðxed value, gAW
B
general thermodynamic equation for the Helmholtz free would allow the Ñuid, at least at the temperature 300 K, to
energy, F \ U [ T S, is applied, the entropy for a hard-core possess a negative entropy within the range of pressures
gas is given explicitly by the CarnahanÈStarling extension of between that determined by the sign change of the entropy
SPT as : and gAW.

SSPT \ Nk
A 5
] ln
A V B [
B
g(4 [ 3g)
\ SIG ] Shc. (8)
The inadequacies of the local stability conditions set out in
eqn. (1) to predict the AlderÈWainwright transition may be
B 2 nN j3 (1 [ g)2
A noted by calculation of the isothermal compressibility,
If the last term in eqn. (8) is neglected, it becomes SakurÈ [V (dp/dV ) , the isometric thermal variance of the pressure,
T
Tetrode equation, to which the SPT equation for entropy (dp/dT ) , the isobaric thermal variance of the volume,
V
reduces in circumstances of dilute densities. (dV /dT ) , and the isobaric speciÐc heat, C , at the density at
p p
The two leading terms in the entropy, the ideal gas contri- which the system entropy changes sign. All these thermodyna-
bution, SIG, are always positive, and dominate that function in mic variables satisfy the local conditions of stability in the
the region of low and moderate densities. The third term in Ñuid phase at the AlderÈWainwright transition.
the entropy, the hard-core contribution, Shc, is always negative A less widely used determinant of a phase transition is the
and, most importantly, possesses a second-order singularity in isobar inÑection, (dCp/dp) \ 0, the importance of which for
T
the reduced density. Therefore, at sufficiently high densities, investigating the phase stability of Ñuids has been described
the entropy of the hard-core Ñuid must change sign and by Deiters and de Reuck26 and by Bernal,27 who pointed out
become negative. the experimental demonstration of its signiÐcance by Jones
Such behavior by the system would constitute a glaring vio- and Walker.28 When the fourth Clausius equation,
lation of the third law of thermodynamics. Therefore, the (dCp/dp) \ [T (d2V /dT 2) , is applied, the isobar inÑection
T p
density at which the entropy of the hard-core Ñuid changes can be determined from the second isobaric temperature
sign represents an absolute limit for the Ñuid, which cannot derivative of the volume. SPT gives for the second isobaric
exist at that, or higher, densities. This limit determines the temperature derivative of the volume :
AlderÈWainwright, high-density, gasÈsolid phase transition.
Scaled particle theory describes equally the dilute gas and the
dense liquid states. Therefore, the limit determined by SPT for
A B
d2V SPT
\[
V
dT 2 T2
the existence of the system so described must set the boundary p
for a solid phase. This property of the entropy of the hard- (1 [ g)g2(1 ] g ] g2 [ g3)(3 ] 9g [ 6g2 [ 4g3 ] 5g4 [ g5)
core Ñuid is shown graphically in Fig. 1 for a hard-sphere gas, ] .
(1 ] 4g ] 4g2 [ 4g3 ] g4)3
at the temperature 300 K, as a function of its reduced density,
g. The calculated values are represented graphically in Fig. 1 (9)
where are shown also the respective contributions of the ideal
gas and hard-core entropies. Also in Fig. 1 is drawn the verti- The isobar inÑection is determined by the roots of the poly-
cal line which represents the reduced density for the gasÈsolid nomial of eighth degree in the numerator of eqn. (9). Aside
phase transition as estimated by Alder and Wainwright using from its obvious roots at zero and unity, eqn. (9) admits no
the Ðxed value for the reduced-density parameter, real roots for values of the reduced-density parameter, g,

Phys. Chem. Chem. Phys., 1999, 1, 3277È3285 3279


contribution of the kinetic energy of the Ñuid has been
obtained by integrating out the MaxwellÈBoltzmann distribu-
tion of kinetic energies of an ideal gas. Therefore, there exist
restrictions, of both volume and temperature, upon the valid-
ity of the SPT formalism. SpeciÐcally, the SPT partition func-
tion is restricted such that values of the system volume used
satisfy the inequality : V A Nj3. In short, the SPT formalism is
valid only for such volumes for which the average interparticle
separation must be much greater than the particle dimensions.
In Fig. 2, the calculated values of the volume of the hard-
sphere gas for which the entropy vanishes, V (T ; S \ 0), are
shown on a double-logarithmic scale through the temperature
range 0.01È10 000 K, together with the values of its ““ kinetic ÏÏ
volume, V (T ) \ Nj3. The temperature at which V \ V is
kin kin
approximately 0.16 K, and represents an absolute limit for the
validity of the SPT formalism. At such low temperatures, not
only has the system volume become very small, in order to
provide a zero for the entropy equation, eqn. (8), but the
thermal de Broglie wavelength has become relatively much
larger because of the square root of the temperature in its
denominator. The value of temperature at which V \ V is
kin
marked in Fig. 2 by a broad dashed stripe, approximately to
Fig. 2 The volume and temperature limits of the validity of scaled divide visually the regions in which the SPT formalism is valid
particle theory. and invalid. Visual inspection in Fig. 2 of the traces of
V (S \ 0) and V persuades that a reasonable limit of con-
kin
Ðdence for SPT could be placed at about the temperatures
greater than zero or less than one ; of its four real roots, two 5È10 K, for at 10 K, V (S \ 0) B 100V .
are negative and the other two greater than one. The values of kin
The pressures at which the entropy of the Ñuid vanishes,
the second isobaric temperature derivative of the volume, cal- p(S \ 0), are shown in Fig. 3 on a double-logarithmic scale for
culated at the density at which the entropy vanishes for the the range of temperatures 0.1È10 000 K. On Fig. 3, are shown
hard-sphere system, do not vanish. also the values of the pressure for which V \ V , which are
kin
marked with a broad dashed stripe, again to indicate the pres-
6. The temperature and volume limitations of sure limits of validity of the SPT formalism. The two regions
of validity and invalidity, respectively, are indicated in Fig. 3.
scaled particle theory Shown also on Fig. 3 are the pressures estimated for the phase
The pressures and reduced volumes of the gasÈsolid phase transition by the Ðxed value of the reduced volume of the
transition for a hard-sphere Ñuid have been calculated at tem- AlderÈWainwright transition given by the value of the
peratures from 0.01 to 10 000 K. The hard-sphere Ñuid con- reduced-density parameter gAW \ 0.67114. As may be noted in
sidered is characterized by a hard-core covolume of 20 cm3 Fig. 3, the predictions of the gasÈsolid transition pressures
mol~1 and a mass of 35 u ; these parameters are typical of from these two calculations give very similar values over a
light gases such as methane or nitrogen. For comparison, the range of temperatures extending three orders of magnitude. At
principal results of the investigation have been carried out high temperatures, the estimate obtained from the Ðxed value
also for a gas with covolume and mass Ðve times greater. of gAW underestimates the transition pressure. At low tem-
SPT uses a factored partition function, eqn. (4), in which the peratures, the estimates of the transition pressure are plainly
too small, and certainly incorrect ; for at temperatures less
than approximately 500 K, a phase transition which might
occur at the pressures estimated by gAW would allow the Ñuid
to exist with negative entropy. This error is shown for the
lower temperatures and shows clearly the errors developed by
the estimate of a Ðxed, temperature-independent, value for the
reduced-density parameter at the gasÈliquid phase transition.
For temperatures less than 100 K, the assumption that the
gasÈsolid phase transition should occur at gAW \ 0.67114
would allow for the Ñuid system to exist in a state of negative
entropy over ranges of almost an order of magnitude in pres-
sure.

7. The high-pressure, gas–solid Alder–Wainwright


transition and its temperature dependence
In Fig. 4, are shown the values of the reduced-density param-
eter at the transition point where the entropy vanishes,
g(S \ 0), as a function of temperature over the range 10È
10 000 K. Shown also on Fig. 4 is the trace of the Ðxed value
of the reduced-density parameter estimated for the transition
point by Alder and Wainwright, gAW \ 0.67114. Although the
Fig. 3 The pressure and temperature dependence limits of the valid- ordinate has been drawn with a very expanded, linear scale in
ity of scaled particle theory, and the pressures of the AlderÈ Fig. 4, the value of the reduced-density parameter at the tran-
Wainwright transition, as functions of temperature, p(S \ 0 ; T ). sition point determined by the vanishing of the entropy di†ers

3280 Phys. Chem. Chem. Phys., 1999, 1, 3277È3285


a gasÈsolid phase transition as its temperature approaches
absolute zero. In Fig. 5, the pressures of the gasÈsolid phase
transition may be observed as a function of temperature over
the low range 0.1È15 K on a linear scale. While taking cogni-
zance of the temperature limitations of the SPT formalism,
speciÐcally that the details of the description of the system are
unreliable at temperatures less than approximately 1È5 K, the
values of the pressure at which the entropy of the Ñuid
becomes zero are nonetheless seen in Fig. 5 to decrease
rapidly with temperature and to approach zero as T ] 0.
Between the temperatures 5È15 K, the values of the pressure
at which the Ñuid system becomes absolutely unstable against
transition to the solid phase decrease by approximately half
an order of magnitude, while the temperature decreases by
only 10 degrees. Thus, p(S \ 0) ] 0 as T ] 0, and faster than
does the temperature itself. The shape of the curve which rep-
resents the pressure of the AlderÈWainwright transition at low
temperature argues strongly that, as the temperature of the
hard-sphere gas approaches absolute zero, the Ñuid becomes
increasingly unstable at any pressure. The pressure deter-
mined by gAW is also drawn on Fig. 5, and is easily seen to
represent unphysical behavior at low temperatures.
Fig. 4 Reduced-density parameter, g, at the transition point deter-
mined by S \ 0, as a function of temperature.

9. The entropy of the gas–solid phase


by that estimated at 0.67114 by only a few percent : at 100 K, transformation estimated by SPT
gS/0 \ 0.63206, 5.8% lower than gAW ; at 1000 K, It has been shown that a hard-sphere system in the Ñuid state,
gS/0 \ 0.68227, 1.6% higher. as described by SPT, is characterized by limits determined by
the stability condition imposed by the third law of thermody-
namics. That such limit determines a ÑuidÈsolid separation is
8. The temperature dependence predicted by SPT easily recognized. The limit set by the change of sign of the
for the Alder–Wainwright transition at low system entropy occurs at densities always greater by approx-
temperatures imately 30% than that of the close-packed conÐguration. Cer-
tainly the hard-sphere system exists at such densities, as well
A consequence of the third law requires that the speciÐc heat, as at greater until that of the close-packed limit. Because SPT
however measured, must vanish at T ] 0 at least as fast as describes the Ñuid system equally in the dilute gas or dense
T (1`d) for d [ 0. However, as has been noted in Section 4, the liquid circumstances, the stability limit must be that for the
speciÐc heat for a hard-core Ñuid described by SPT remains at separation of the system in its ÑuidÈsolid phases.
all temperatures simply that of an ideal gas : CSPT \ 3/2Nk . Because SPT predicts a solidÈÑuid phase transition for a
V B
By such property, SPT describes a Ñuid phase which fails to hard-sphere gas, it is of interest to calculate the latent heat of
manifest correct physical behavior at very low temperatures ; a transformation at its sublimation or melting point. Because
fact which suggests that SPT might indicate the presence of a the latent heat of transition represents the enthalpic di†erence
low-temperature gasÈsolid phase transition. between two phases, the entropy of transformation cannot be
SPT does indeed strongly indicate that the Ñuid undergoes calculated presently for the hard-sphere system because an
exact thermodynamic description is not yet available for the
hard-sphere solid. Here follows an approximate calculation of
the limiting value of the SPT entropy of melting at low tem-
peratures.
As a hard-sphere system approaches absolute zero, its latent
heat of transition for the solidÈÑuid phase transition, *L sf,
approaches a constant :
*L (T )sf \ *Hsf \ (U(T )fluidU(T )solid)
] psf(V (T )fluid [ V (T )solid), (10)
in which psf represents the pressure of the solidÈÑuid phase
transition. As shown in Fig. 5 of the previous section, the pres-
sure of the solidÈÑuid phase transition, psf, goes rapidly to
zero at low temperatures. Therefore, although there exists a
volume di†erence between the two phases, the second term in
parentheses on the right side of eqn. (10) decreases rapidly as
the system temperature approaches absolute zero. The Ðrst
term in parentheses on the right side of eqn. (10) does not
vanish as the system temperature approaches absolute zero.
The internal energy of the gas phase vanishes linearly with
temperature,
U(T )fluid \ 3 Nk T , (11)
Fig. 5 Low temperature behavior of the AlderÈWainwright tran- 2 B
sition predicted by S \ 0. while that of the solid may be assumed to have the general

Phys. Chem. Chem. Phys., 1999, 1, 3277È3285 3281


form sphere solid or its melting entropy at general temperatures. As

U(T )solid D 3 Nk T
C A B
2p4 T 3 D
] É É É O(exp([T D/T )) , (12)
has been pointed out in the previous section, the contribution
of the low-temperature limit of the melting entropy is approx-
2 B 5 TD imately 3 cal mol~1 K~1, and this value is of the same order
in which T D \ +wD/k represents the e†ective Debye tem- as, and very closely approximates, the entire entropy of
B melting in simple, real materials. Therefore, in order to
perature ; and wD \ kDc where kD represents the geometric,
s examine how inclusion of the entropy of melting might alter
maximal, Debye wave number and c the velocity of sound.
s the details of the AlderÈWainwright transition, as determined
Thus the limit of the entropic change in passing from the
solid to gas phase is : by SPT, the pressures for that transition have been calculated

C
lim (*Ssf)SPT \
D
*L (T )sf
ÈÈÈÕ 3 Nk
using the condition S(p, T ) ] (*S)sf \ 0. In Fig. 6, are shown
the pressures of the AlderÈWainwright transition as deter-
T sf 2 B mined by the vanishing of the system entropy as calculated
T?0 both from the Ñuid phase alone, and also with inclusion of the
B 2.98 cal mol~1 K~1. (13) contribution of the entropy of melting, as estimated from its
This interesting property has not been previously reported low-temperature limit. The data represented in Fig. 6, show
in the extensive literature dealing with SPT. The entropy of clearly that the inclusion of the entropy of melting has no
fusion at the melting point for almost all metals lies in the e†ect upon the qualitative features of the AlderÈWainwright
range 1.7È2.6 cal mol~1 K~1.29 That the entropy of melting transition and produces only very small quantitative change
for real metals should be less than that given by the SPT limit, above the temperature of approximately 100 K. Furthermore,
eqn. (13), should be expected, for the limit to which eqn. (13) the low-temperature behavior of the AlderÈWainwright tran-
refers is that for which the temperature vanishes. Real metals sition as predicted by SPT, characterized by the vanishing of
fuse at high temperatures, at which the calculated entropy the Ñuid phase at all pressures as the temperature goes to
would have subtracted from eqn. (12) contributions for both zero, remains. The magnitude of the calculated reduced-
the binding energy of the metal and also its volume change at density parameter at the density of the AlderÈWainwright
melting. The typically high melting temperature of most transition as predicted by SPT, g(S ] (*S)sf \ 0), deserves par-
metals indicates substantial binding energy, and thereby a sig- ticular attention : In Fig. 6, it may be observed that the
niÐcant contribution to the change in enthalpy from the term reduced-density parameter at the AlderÈWainwright tran-
*Usl/T . An example of a real material composed of volatile sition is essentially indistinguishable from the Ðxed value,
molecules of almost spherical shape, the nonassociating gAW \ 0.67114, estimated by Alder and Wainwright, over a
hydrocarbon tetramethylmethane (2-2-dimethylpropane), range of temperatures extending almost three orders of magni-
C H , melts at 250.2 K and possesses a latent heat tude.
5 12 Because any density at which the system entropy vanishes
*Hsl \ 778 cal mol~1, such that its entropy of transformation
*Hsl/T sl \ 3.0 cal mol~1 K~1.30 Thus, SPT predicts a limit- determines a point of absolute instability for that phase, the
ing value for the transformation entropy which is close to the densities of the AlderÈWainwright transition calculated by
measured values of real materials. SPT should be considered the solid-end points of the gasÈ
solid coexistence line for the phase transition, at each tem-
10. Discussion perature. The positions of the gas-end of the gasÈsolid
coexistence line can be determined only with an exact solution
The purpose of this short paper has been to demonstrate that of the statistical mechanics of the hard-sphere solid, which
SPT accurately predicts the high-density, gasÈsolid AlderÈ does not exist.
Wainwright transition in a hard-sphere system. The qualit- The description of the high-density AlderÈWainwright tran-
ative features of the mathematical computations described sition given above by SPT is not restricted to artiÐcial hard-
above remove all doubts on that question. Furthermore, these sphere gases ; it holds equally true in all its qualitative features
results show that SPT additionally describes interesting fea- for real Ñuids. The mathematical description of the entropy of
tures of the hard-sphere system not previously been reported : every real Ñuid system must possess a component which
the temperature dependence of the reduced-density parameter accounts for the entropy of exclusion, as determined by SPT.
at the gasÈsolid phase transition, g(T ) ; and the require-
S/0
ment which SPT appears to place upon the hard-core system,
as the system temperature approaches absolute zero, to con-
dense into the solid phase at Ðnite temperatures and pressures.
The temperature dependence of the density at which the
hard-sphere system undergoes a ÑuidÈsolid phase transition,
g(T ) , should be expected. The phase rule of Gibbs in no
S/0
way depends upon any detail of the intermolecular potential,
and holds true for a hard-sphere system as any other. A
single-component, hard-sphere Ñuid in a single phase must be
described by two independent, intensive variables, and at its
binary-phase transition, one.
Although it has been shown that the Ñuid phase of a hard-
core system, as described by SPT, is characterized by limits
determined by the stability condition imposed by the third
law of thermodynamics, the detailed quantities calculated
above cannot be considered to be strictly accurate, for such
were calculated by considering only the entropy of the gas
state. An exact calculation of the pressures of the AlderÈ
Wainwright transition must include, in addition to the
entropy of the gas phase, that of the solid phase of the hard-
core system. However, the thermodynamics of the hard-sphere
solid have not yet been worked out, and it is therefore not Fig. 6 Pressure of the AlderÈWainwright transition, determined with
possible to calculate exactly either the entropy of the hard- inclusion of the entropy of transformation, (*S)gf B 3/2k .
B
3282 Phys. Chem. Chem. Phys., 1999, 1, 3277È3285
The entropy of exclusion is obtained from the impenetrable given by the Tonks equation :41
property of real molecules, and is a consequence of the mathe-
matics of statistical geometry, and is always negative. Further- pSPT(1vdim) \
Nk T
B
1
\ Nk T
1
]
Cg D
more, the negative term which describes the entropy of L (1 [ g) B L L (1 [ g)
exclusion of a Ñuid will possess always the second-order singu- \ pIG(1vdim) ] phc(1vdim) (14)
larity determined by SPT, Shc(g) D [h(g)/(1 [ g)2, where h(g)
represents any non-singular function of g, usually a poly- where g \ Np/L , in which p represents the length of the hard
nomial. Even when the complications of a non-spherical shape rod and L the system dimension (““ volume ÏÏ). The Helmholtz
of the hard-core have been taken into account, as has been free energy of the one-dimensional hard-rod system is given
done by, for example, Boublik et al.,31 Chapman et al.,32 explicitly as :
Nezbeda et al.,33 Vimalchand et al.,34 and Zhou et al.,35 the
second-order singularity in the reduced density remains. The FSPT(1vdim) \ [ Nk T ln
L CA B
] 1 [ ln(1 [ g)
D
same fact holds also when the temperature dependence of the B Nj
hard core is introduced, as has been done by, for example, \ FIG(1vdim) ] Fhc(1vdim). (15)
Deiters et al.36,37 The entropy of every real gas system is posi-
tive at low and modest pressures at all temperatures, of The hard-core terms in the free energy, and the canonical par-
course, for if not, the gas state would always be unstable tition function, of a hard-rod system in one dimension are
against transition into the solid state and could not exist. purely entropic, as for the three-dimensional case ; and the
However, as stated above, the total entropy of every real gas internal energy, U, of the hard-rod system is exactly 1/2Nk T .
B
system possesses also, at all temperatures, a negative com- The entropy of the hard-rod system, given explicitly by
ponent, the entropy of exclusion attributable to the universal S \ (U [ F)/T , is :
property of impenetrability of its molecules as described by
SPT. The (negative) component, the entropy of exclusion, by SSPT(1vdim) \ Nk
3C] ln
L A B
] ln(1 [ g) .
D (16)
B 2 Nj
virtue of its second-order singularity, is unbounded at high-
densities. Therefore, at some sufficiently high density, the Ñuid The kinetic volume, V , of a D-dimensional system may be
phase of every real gas must become absolutely unstable. That kin
written as (V )Dvdim \ (Nj3)D@3. The condition of validity for
instability is the famous AlderÈWainwright transition. The kin
the hard-core Ñuid in any dimension, D, can be written as
AlderÈWainwright transition is not an artifact of the machine- (Nj3)D@3/LD \ V /V @ 1.
calculated simulations with which Alder and Wainwright Ðrst kin
If the one-dimensional hard-rod system were compressed to
determined the high-density gasÈsolid phase transition ; it is a the length at which the entropy, as given by eqn. (16), would
real phenomenon which occurs in all real Ñuids. vanish, its length would then have to satisfy the equation :
Because there are very few models in quantum statistical
mechanics which admit exact solutions, the constraints of the V AB
kin \ exp
3
(1 [ g) B 4.482(1 [ g). (17)
third law have not often been applied for stability analysis. V 2
S/0
Therefore, and in order to show that such analysis does not However, both eqn. (17) and the condition of validity cannot
predict a spurious instability or phase transition, in two be satisÐed simultaneously. Because the limits of the reduced-
Appendices which follow, the identical analysis as has been density parameter are 0 O g O 1, the value of the system
used for the three-dimensional hard-sphere gas is applied to, length at which the expression for the entropy, as given by
respectively, the one-dimensional hard-body, or ““ hard rod ÏÏ, eqn. (16), vanishes, L , clearly lies outside the limits of
system, and the two-dimensional, or ““ hard-disk ÏÏ, system. In S/0
validity of the equation of state, and of the hard-rod partition
those Appendices is shown that, the constraints of the third function from which it was obtained. (The mathematically
law, when applied to SPT and the one- and two-dimensional valid but physically meaningless values of, say, g \ 0.99999
hard-core systems, do not predict a spurious instability for the . . . , are not considered, for such must be interpreted simply as
Ðrst case, and do predict one-such, and its accompanying Ðrst- the system compressed to close-packing, not as a phase tran-
order, ÑuidÈsolid phase transition, in the second case, exactly sition. For all ÑuidÈsolid phase transitions, the change in
as demonstrated by molecular dynamics computer simula- density at the temperature of transformation is approximately
tions.38h40 10È15%. Therefore, when considering eqn. (17), it would be
Notwithstanding its use of the modern formalism of SPT, reasonable to restrict the reduced-density parameter such that
this prediction of the AlderÈWainwright, ÑuidÈsolid phase g O D0.9.)
transition, above, is very much in the spirit of the analysis of Therefore, the stability condition that the entropy of the
the gasÈliquid phase transition by van der Waals. The tradi- system must always be positive does not predict any phase
tional prediction of the low-temperature gasÈliquid phase transition for the one-dimensional hard-core system at any
transition obtained by the van der Waals equation of state volume within the limits of validity of its description. Applica-
does not strictly describe any liquid conÐguration. When the tion of the constraints of the third law to the one-dimensional
van der Waals equation of state is used to describe a gas, its system does not involve any prediction of instability or phase
isothermal compressibility is observed to change sign at a transition for the one-dimensional hard-rod gas, and is consis-
certain (critical) temperature ; and for all lower temperatures, tent with other stability analyses.
the gasÈliquid phase coexistence line is inferred a fortiori,
using the (appropriately named) ““ Maxwell construction ÏÏ.
Similarly, when SPT is used to describe a hard-sphere gas, its Appendix B Application of the constraints of the
entropy is observed to change sign at a certain density ; and third law to the two-dimensional, hard-disk system :
for all higher densities, the solid state is inferred a fortiori. SPT develops the exact solution also for the equation of state
of the two-dimensional hard-body system, or ““ hard-disk ÏÏ
system :
Appendix A Application of the constraints of the
Nk T 1 1 C
g(2 [ g)
third law to the one-dimensional hard-rod system
The statistical mechanical problem of a one-dimensional hard-
pSPT(2vdim) \ B
A (1 [ g)2
\ Nk T
B A
]
A(1 [ g)2
D
body gas admits an exact solution, and no phase transition.
The equation of state of a one-dimensional hard-body Ñuid is \ pIG(2vdim) ] phc(2vdim) (18)

Phys. Chem. Chem. Phys., 1999, 1, 3277È3285 3283


where g \ opp2/4, in which p represents the diameter of the
hard disk and A the system area (““ volume ÏÏ). The Helmholtz
free energy of the two-dimensional hard-disk system is given
explicitly as :
C A B
FSPT(2vdim) \ Nk T [ ln
V
[1]
g
\ ln(1 [ g) ;
D
B V (1 [ g)
kin
(19)
and the entropy of the hard-disk system is therefore given
explicitly by S \ (U [ F)/T as :
C
SSPT(2vdim) \ Nk 2 ] ln
A B
V
[
g D
] ln(1 [ g) . (20)
B V (1 [ g)
kin
If the two-dimensional hard-disk system were compressed
to the area at which the entropy, as given by eqn. (20), would
vanish, its area would then have to satisfy the equation :
V
kin \ y(g) \ (1 [ g)exp 2 [
A g B
V (1 [ g)
S/0
B 7.39(1 [ g)exp [
A g
.
B (21)
Fig. 8 Area of the two-dimensional hard-disk system at the point
S \ 0, and its kinetic area, A \ (Nj3)2@3, as functions of tem-
(1 [ g) kin
pereature.
At low densities, eqn. (21) and the condition of validity obvi-
ously cannot be satisÐed simultaneously. However, the limits
of the reduced-density parameter, or reduced density, are
0 O g O 1, and the value of the function y(g) diminishes
rapidly for values of its argument greater than approximately
6. A plot of y(g) is shown in Fig. 7 for values of the argument
between 10~4È0.95, in which can be seen that the require-
ments of inequality of Fig. 7 are easily satisÐed for reduced
densities greater than 6. The line V /V \ 1, which marks the
kin
absolute limits of validity of the SPT formalism is also shown
as a broad, dashed gray band on Fig. 7.
The region within which the gasÈsolid phase transition
occurs is to the right of the trace of y(g) and (well) below the
line A /A \ 1. The limits of the validity of the SPT formal-
kin
ism in two dimensions are shown graphically also in Fig. 8 for
a two-dimensional gas of particles characterized by a mass of
35 u and a molar covolume of 20 cm3 mol~1, which values are
typical for a light gas. In Fig. 9 is shown graphically the value
of the reduced-density parameter at its values for which the
system entropy of the Ñuid phase vanishes, g(S \ 0). Similarly
as a real, three-dimensional system, the transition value of the

Fig. 9 Reduced-density parameter at the transition point, g(S \ 0),


of the two-dimensional hard-disk system as a function of temperature.

reduced-density parameter depends weakly upon temperature.


At 100 K, gT \ 0.793 ; at 1000 K, gT \ 0.890 K. For reference,
the value of the reduced-density parameter at close-packing,
gcp \ 2/J3 B 1.15 has also been plotted in Fig. 9.
Therefore, as may be noted in Fig. 9, when the constraints
of the third law of thermodynamics are applied, SPT unam-
biguously establishes an absolute instability of the Ñuid phase
in a two-dimensional system at temperatures greater than 10
K for values of the reduced-density parameter greater than
approximately 0.80È0.88, depending upon the temperature.
Such ÑuidÈsolid phase transition in two-dimensional hard-
disk systems, at exactly these densities, has already been
demonstrated (most recently) by Mitus et al.,38 and (two
decades ago) by Hoover et al.39

References
1 J. F. Kenney, Fluid Phase Equilib., 1998, 148, 21.
2 B. J. Alder and T. E. Wainwright, J. Chem. Phys., 1957, 27, 1208.
Fig. 7 Ratio of kinetic to system area (volume) of a two-dimensional 3 J. G. Kirkwood, J. Chem. Phys., 1935, 3, 300.
hard-disk system at the point S \ 0, as a function of reduced density. 4 M. Born and H. S. Green, Proc. R. Soc. L ondon, 1946, 188, 10.

3284 Phys. Chem. Chem. Phys., 1999, 1, 3277È3285


5 H. Reiss, H. L. Frisch and J. L. Lebowitz, J. Chem. Phys., 1959, 25 B. J. Alder, W. G. Hoover and D. A. Young, J. Chem. Phys., 1968,
31, 369. 49, 3688.
6 H. Reiss, H. L. Frisch, E. Gelfand and J. L. Lebowitz, J. Chem. 26 U. K. Deiters and K. M. de Reuck, International Union of Pure
Phys., 1960, 32, 119. and Applied Chemistry, 1997.
7 H. Reiss, ““Scaled particle theory in the statistical thermodyna- 27 J. D. Bernal, Nature, 1959, 183, 141.
mics of ÑuidsÏÏ, in Adv. Chem. Phys., ed. I. Prigogine, Interscience, 28 G. O. Jones and P. A. Walker, Proc. Phys. Soc. B., 1953, 69, 1348.
New York, 1965, vol. 9. 29 R. H. Fowler and E. A. Guggenheim, Statistical Mechanics, Cam-
8 T. De Donder, L ÏAffinity, Paris, 1936. bridge University Press, Cambridge, 1939.
9 I. Prigogine and R. Defay, Chemical T hermodynamics, Longmans, 30 J. Timmermans and L. De†et, L e Polymorphisme des Composes
London, 1954. Organiques, Gauthier-Villars, Paris, 1939.
10 P. Glansdor† and I. Prigogine, T hermodynamic T heory of Struc- 31 T. Boublik, C. K. Viga and M. Diaz-Pena, J. Chem. Phys., 1990,
ture, Stability, and Fluctuations, John Wiley, New York, 1971. 93, 730.
11 I. Prigogine, Introduction to the T hermodynamics of Irreversible 32 W. G. Chapman, K. E. Gubbins, G. Jackson and M. Radosz, Ind.
Processes, John Wiley, New York, 1967. Eng. Chem. Res., 1990, 29, 1709.
12 I. Prigogine and C. George, Proc. Nat. Acad. Sci., 1983, 80, 4590. 33 I. Nezbeda, Mol. Phys., 1997, 90, 661.
13 H. Reiss, H. L. Frisch and J. L. Lebowitz, ““Mixtures of hard 34 P. Vimalchand, A. Thomas, I. G. Economou and M. D.
spheresÏÏ, in T he Equilibrium T heory of Classical Fluids, ed. H. L. Donohue, Fluid Phase Equilib, 1992, 73, 39.
Frisch and J. L. Lebowitz, W. A. Benjamin, New York, 1964. 35 Y. Zhou, S. W. Smith and C. Hall, Mol. Phys., 1995, 85, 1157.
14 H. Reiss, J. Chem. Phys., 1967, 47, 186. 36 U. K. Deiters and S. L. Randzio, Fluid Phase Equilib, 1995, 103,
15 H. Reiss and D. M. Tully-Smith, J. Chem. Phys., 1971, 55, 1674. 199.
16 H. Reiss and R. V. Casberg, J. Chem. Phys., 1974, 61, 1107. 37 U. K. Deiters, Mol. Phys, 1992, 96, 539.
17 H. Reiss, ““Scaled particle theory of hard sphere Ñuids to 1976ÏÏ, in 38 A. C. Mitus, H. Weber and D. Marx, Phys. Rev., 1997, E55, 6855.
Statistical Mechanics and Statistical Methods in T heory and 39 W. G. Hoover, N. E. Hoover and K. Hansen, J. Chem. Phys.,
Application, ed. U. Landman, Plenum, London, 1977. 1979, 70, 1837.
18 D. M. Tully-Smith and H. Reiss, J. Chem. Phys., 1970, 33, 4015. 40 E. Helfand, H. L. Frisch and J. L. Lebowitz, J. Chem. Phys., 1961,
19 N. F. Carnahan and K. E. Starling, J. Chem. Phys., 1969, 51, 635. 34, 1037.
20 H. F. Baker, Proc. L ondon Math. Soc., 1902, 34, 347. 41 L. Tonks, Phys. Rev., 1936, 50, 955.
21 J. E. Campbell, Proc. L ondon Math. Soc., 1898, 29, 14.
22 H. F. Baker, Proc. L ondon Math. Soc., 1903, 2, 293.
23 H. F. Baker, Proc. L ondon Math. Soc., 1904, 3, 24.
24 F. Hausdor†, Ber. V erh. Saechs. Akad. W iss. L eipzig, Math-
Naturwiss., 1906, 58, 19. Paper 9/01700C

Phys. Chem. Chem. Phys., 1999, 1, 3277È3285 3285

Das könnte Ihnen auch gefallen