equation of state of a multicomponent d-dimensional hard-sphere fluid

5
arXiv:cond-mat/0204246 v1 11 Apr 2002 Equation of state of a multicomponent d-dimensional hard-sphere fluid [Mol. Phys. 96, 1–5 (1999)] Andr´ es Santos, 1 Santos Bravo Yuste, 1 and Mariano L´ opez de Haro 2 1 Departamento de F´ ısica, Universidad de Extremadura, E-06071 Badajoz, Spain 2 Centro de Investigaci´on en Energ´ ıa, U.N.A.M., Apartado Postal 34, Temixco, Mor. 62580, Mexico A simple recipe to derive the compressibility factor of a multicomponent mixture of d-dimensional additive hard spheres in terms of that of the one-component system is proposed. The recipe is based (i) on an exact condition that has to be satisfied in the special limit where one of the components corresponds to point particles; and (ii) on the form of the radial distribution functions at contact as obtained from the Percus-Yevick equation in the three-dimensional system. The proposal is examined for hard discs and hard spheres by comparison with well-known equations of state for these systems and with simulation data. In the special case of d = 3, our extension to mixtures of the Carnahan-Starling equation of state yields a better agreement with simulation than the already accurate Boubl´ ık-Mansoori-Carnahan-Starling-Leland equation of state. Due to their importance in liquid state theory, for years researchers have proposed empirical or semi-empirical (analytical) equations of state of various degrees of com- plexity for one-component hard-sphere fluids. Notable among these, is the celebrated Carnahan-Starling (CS) equation of state [1], which is not only rather simple but also accurate in comparison with computer simulation data. In the case of hard-sphere mixtures, the proposals, also empirical or semi-empirical in nature, are much more limited, with the Boubl´ ık-Mansoori-Carnahan-Starling- Leland (BMCSL) equation [2] standing out as the usual favorite. The situation for one-component hard-disc flu- ids is rather similar. Here, no analog of the CS equation using the 1 3 (v) + 2 3 (c) recipe has been derived, due to the absence of an analytical solution of the Percus-Yevick (PY) equation in this instance. Nevertheless, accurate and simple equations of state have been proposed, such as the popular Henderson equation [3] and the recent one by the present authors [4]. Hard-disc mixtures, on the other hand, have received much less attention and the proposed equations of state for these systems are rather scarce. Given this scenario, the major aim of this paper is to show that, on the basis of a simple recipe, accurate equations of state of a multicomponent d-dimensional ad- ditive hard-sphere mixture may be derived, requiring the equation of state of the one-component system as the only input. The recipe makes use of a consistency con- dition that arises in the case that one of the components in the mixture has a vanishing size, as well as from some insight gained from the form of the radial distribution functions at contact given by the solution of the Percus- Yevick equation for a hard-sphere fluid in three dimen- sions [5]. Let us consider an N -component system of hard spheres in d dimensions. The total number density is ρ, the set of molar fractions is {x 1 ,...,x N }, and the set of diameters is {σ 1 ,...,σ N }. The volume packing frac- tion is η = v d ρσ d , where v d =(π/4) d/2 /Γ(1 + d/2) is the volume of a d-dimensional sphere of unit diameter and σ n 〉≡ i x i σ n i . In the case of a polydisperse mix- ture (N →∞) characterized by a size distribution f (σ), σ n 〉≡ dσσ n f (σ). Our goal is to propose a simple equation of state (EOS) for the mixture, Z (N) (η), con- sistent with a given EOS for a one-component system, Z (1) (η), where Z = p/ρk B T is the usual compressibil- ity factor. A consistency condition appears when one of the species, say the N -th, has a vanishing diameter, i.e. σ N 0. In that case, Z (N) (η) (1 x N )Z (N-1) (η)+ x N 1 η . (1) At a more fundamental level, we will consider the contact values of the radial distribution functions, g (N) ij (σ ij ), the knowledge of which implies that of the EOS through the relation Z (N) (η)=1+2 d-1 η N i=1 N j=1 x i x j σ d ij σ d g (N) ij (σ ij ). (2) Taking as a guide the form of g (N) ij obtained through the solution of the PY equation for hard spheres (d = 3) [5], we propose to approximate g (N) ij by a linear interpo- lation between g (1) and (1 η) -1 , namely g (N) ij (σ ij )= 1 1 η + g (1) (σ) 1 1 η σ d-1 σ d σ i σ j σ ij . (3) When the above ansatz is inserted into equation (2), one gets Z (N) (η) 1= Z (1) (η) 1 2 1-d Δ 0 + η 1 η 1 Δ 0 + 1 2 Δ 1 , (4) where Δ p = σ d+p-1 σ d 2 d-1 n=p (d + p 1)! n!(d + p 1 n)! σ n-p+1 〉〈σ d-n , (p =0, 1). (5)

Upload: ervar-potro-chavez

Post on 25-Dec-2015

3 views

Category:

Documents


1 download

DESCRIPTION

Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

TRANSCRIPT

Page 1: Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

arX

iv:c

ond-

mat

/020

4246

v1

11

Apr

200

2Equation of state of a multicomponent d-dimensional hard-sphere fluid

[Mol. Phys. 96, 1–5 (1999)]

Andres Santos,1 Santos Bravo Yuste,1 and Mariano Lopez de Haro2

1Departamento de Fısica, Universidad de Extremadura, E-06071 Badajoz, Spain2Centro de Investigacion en Energıa, U.N.A.M.,

Apartado Postal 34, Temixco, Mor. 62580, Mexico

A simple recipe to derive the compressibility factor of a multicomponent mixture of d-dimensionaladditive hard spheres in terms of that of the one-component system is proposed. The recipe is based(i) on an exact condition that has to be satisfied in the special limit where one of the componentscorresponds to point particles; and (ii) on the form of the radial distribution functions at contactas obtained from the Percus-Yevick equation in the three-dimensional system. The proposal isexamined for hard discs and hard spheres by comparison with well-known equations of state forthese systems and with simulation data. In the special case of d = 3, our extension to mixtures ofthe Carnahan-Starling equation of state yields a better agreement with simulation than the alreadyaccurate Boublık-Mansoori-Carnahan-Starling-Leland equation of state.

Due to their importance in liquid state theory, for yearsresearchers have proposed empirical or semi-empirical(analytical) equations of state of various degrees of com-plexity for one-component hard-sphere fluids. Notableamong these, is the celebrated Carnahan-Starling (CS)equation of state [1], which is not only rather simple butalso accurate in comparison with computer simulationdata. In the case of hard-sphere mixtures, the proposals,also empirical or semi-empirical in nature, are much morelimited, with the Boublık-Mansoori-Carnahan-Starling-Leland (BMCSL) equation [2] standing out as the usualfavorite. The situation for one-component hard-disc flu-ids is rather similar. Here, no analog of the CS equationusing the 1

3 (v)+ 23 (c) recipe has been derived, due to the

absence of an analytical solution of the Percus-Yevick(PY) equation in this instance. Nevertheless, accurateand simple equations of state have been proposed, suchas the popular Henderson equation [3] and the recentone by the present authors [4]. Hard-disc mixtures, onthe other hand, have received much less attention and theproposed equations of state for these systems are ratherscarce. Given this scenario, the major aim of this paperis to show that, on the basis of a simple recipe, accurateequations of state of a multicomponent d-dimensional ad-ditive hard-sphere mixture may be derived, requiring theequation of state of the one-component system as theonly input. The recipe makes use of a consistency con-dition that arises in the case that one of the componentsin the mixture has a vanishing size, as well as from someinsight gained from the form of the radial distributionfunctions at contact given by the solution of the Percus-Yevick equation for a hard-sphere fluid in three dimen-sions [5].

Let us consider an N -component system of hardspheres in d dimensions. The total number density isρ, the set of molar fractions is {x1, . . . , xN}, and the setof diameters is {σ1, . . . , σN}. The volume packing frac-tion is η = vdρ〈σd〉, where vd = (π/4)d/2/Γ(1 + d/2) isthe volume of a d-dimensional sphere of unit diameterand 〈σn〉 ≡ ∑

i xiσni . In the case of a polydisperse mix-

ture (N → ∞) characterized by a size distribution f(σ),

〈σn〉 ≡∫

dσσnf(σ). Our goal is to propose a simple

equation of state (EOS) for the mixture, Z(N)(η), con-sistent with a given EOS for a one-component system,Z(1)(η), where Z = p/ρkBT is the usual compressibil-ity factor. A consistency condition appears when one ofthe species, say the N -th, has a vanishing diameter, i.e.σN → 0. In that case,

Z(N)(η) → (1 − xN )Z(N−1)(η) +xN

1 − η. (1)

At a more fundamental level, we will consider the contact

values of the radial distribution functions, g(N)ij (σij), the

knowledge of which implies that of the EOS through therelation

Z(N)(η) = 1 + 2d−1ηN

i=1

N∑

j=1

xixj

σdij

〈σd〉g(N)ij (σij). (2)

Taking as a guide the form of g(N)ij obtained through

the solution of the PY equation for hard spheres (d = 3)

[5], we propose to approximate g(N)ij by a linear interpo-

lation between g(1) and (1 − η)−1, namely

g(N)ij (σij) =

1

1 − η+

[

g(1)(σ) − 1

1 − η

] 〈σd−1〉〈σd〉

σiσj

σij. (3)

When the above ansatz is inserted into equation (2), onegets

Z(N)(η) − 1 =[

Z(1)(η) − 1]

21−d∆0

1 − η

(

1 − ∆0 +1

2∆1

)

, (4)

where

∆p =〈σd+p−1〉〈σd〉2

d−1∑

n=p

(d + p − 1)!

n!(d + p − 1 − n)!〈σn−p+1〉〈σd−n〉,

(p = 0, 1). (5)

Page 2: Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

2

This form of the EOS complies with the requirement (1).Note that equation (4) expresses Z(N)(η) − 1 as a lin-ear combination of Z(1)(η) − 1 and (1 − η)−1 − 1, butthe dependence of the coefficients on the size distribu-tion is much more involved than in equation (3). Thekey outcome of this paper is the EOS given by equation(4), in which the compressibility factor of the mixtureis obtained from that of the one-component system forarbitrary values of the dimensionality d and the num-ber of components N . It is worth noticing that in theone-dimensional case, equation (4) yields the exact re-sult Z(N)(η) = Z(1)(η).

As a straightforward application of equa-

tion (4), one can easily get B(N)n =

vn−1d 〈σd〉n−1

[

21−d∆0b(1)n + 1 − ∆0 + 1

2∆1

]

, where

the virial coefficients B(N)n are defined by

Z(N)(η) = 1 +∑

n=2 B(N)n ρn−1. and where b

(1)n

are the reduced virial coefficients of the one-component

system, i.e. Z(1)(η) = 1 +∑

n=2 b(1)n ηn−1.

We will now focus on the case of hard discs (d = 2).Equation (4) then becomes

Z(N)(η) = Z(1)(η)〈σ〉2〈σ2〉 +

1

1 − η

(

1 − 〈σ〉2〈σ2〉

)

. (6)

The relationship between Z(N)(η) and Z(1)(η) as givenby equation (6) rests on a different rationale from thatpertaining to another simple proposal, namely the Con-formal Solution Theory (CST) [6, 7]. In this latter the-

ory, the EOS reads Z(N)CST(η) = Z(1)(ηeff) with ηeff =

12

(

1 + 〈σ〉2/〈σ2〉)

η, but we note that this equation does

not comply with the general requirement (1). If the one-component system is assumed to be described by theScaled Particle Theory (SPT) [8], our extension to mix-tures takes on a particularly simple form:

Z(N)SPT(η) =

1 −(

1 − 〈σ〉2/〈σ2〉)

η

(1 − η)2. (7)

This is precisely the true SPT EOS for mixtures [7],which is indeed rewarding. The EOS Z(1)(η) =[

1 − 2η + (2η0 − 1)η2/η20

]

−1, where η0 =

√3π/6 is the

crystalline close-packing fraction, has been recently pro-posed by us [4] to describe a one-component system.When this EOS, hereafter referred to as the SHY EOSfollowing the nomenclature introduced in reference [9],is substituted into equation (6), we obtain the followingextension:

Z(N)eSHY(η) =

〈σ〉2/〈σ2〉1 − 2η + (2η0 − 1)η2/η2

0

+1

1 − η

(

1 − 〈σ〉2〈σ2〉

)

.

(8)The well-known Henderson (H) EOS [3] can also be ex-tended:

Z(N)eH (η) =

1 −(

1 − 〈σ〉2/〈σ2〉)

η + (b(1)3 − 3)

(

〈σ〉2/〈σ2〉)

η2

(1 − η)2,

(9)

where b(1)3 = 16

3 − 4π

√3. This equation is quite similar

to the one proposed by Barrat et al. [7], the only dif-ference being that the coefficient of η2 in the numerator

is b(N)3 − 1 − 2〈σ〉2/〈σ2〉, where b

(N)3 is the exact (re-

duced) third virial coefficient. Nevertheless, since b(N)3 is

well approximated by 1 + (b(1)3 − 1)〈σ〉2/〈σ2〉, both EOS

are practically indistinguishable. Although we could con-sider the extensions of other EOS originally proposed fora one-component system of hard discs (for a list of manysuch EOS we refer the reader to references [4, 9]), forthe sake of simplicity we will restrict our analysis to theSPT, eSHY and eH EOS.

Let us now consider the virial coefficients B(N)n for

hard discs. It follows that in this case B(N)n =

(π/4)n−1 〈σ2〉n−1[

1 + (b(1)n − 1)〈σ〉2/〈σ2〉

]

. This equa-

tion yields the exact second virial coefficient [7], thehigher coefficients being approximate. In the par-ticular case of a binary mixture, the composition-

independent coefficients B(2)n1,n2

are defined through

B(2)n =

∑nn1=0

n!n1!(n−n1)!

B(2)n1,n−n1

xn1

1 xn−n1

2 . According

to equation (6),

B(2)n1,n2

=(π

4

)n−1

σ2(n−1)1 α2(n2−1)

[n1

nα2 +

n2

n+

(

b(1)n − 1

)

×(

n1(n1 − 1)

n(n − 1)α2 +

2n1n2

n(n − 1)α +

n2(n2 − 1)

n(n − 1)

)]

,

(10)

where n = n1 + n2 and α = σ2/σ1. This form has thesame structure as the interpolation formula suggested byWheatley [10]. In fact, he proposes an EOS (henceforthlabelled as W) of the form

Z(2)W (η) =

∑7n=0 cnηn

(η − η0)2, (11)

where the coefficients cn are chosen so as to reproducethe first eight virial coefficients given by the interpolationformula [10].

Now, let us consider the case d = 3. Equation (4) thenyields

Z(N)(η) = = 1 +[

Z(1)(η) − 1] 〈σ2〉

2〈σ3〉2(

〈σ2〉2 + 〈σ〉〈σ3〉)

1 − η

[

1 − 〈σ2〉〈σ3〉2

(

2〈σ2〉2 − 〈σ〉〈σ3〉)

]

. (12)

Using the CS EOS [1], Z(1)(η) = (1+η+η2−η3)/(1−η)3,the result may be expressed as

Z(N)eCS(η) = Z

(N)BMCSL(η)+

η3

(1 − η)3〈σ2〉〈σ3〉2

(

〈σ〉〈σ3〉 − 〈σ2〉2)

,

(13)where the BMCSL EOS [2] is

Z(N)BMCSL(η) =

1

1 − η+

(1 − η)2〈σ〉〈σ2〉〈σ3〉 +

η2(3 − η)

(1 − η)3〈σ2〉3〈σ3〉2 .

(14)

Page 3: Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

3

As another, example, let us consider the Carnahan-Starling-Kolafa (CSK) EOS [11], Z(1)(η) = [1 + η + η2 −2η3(1 + η)/3]/(1 − η)3. Its extension is

Z(N)eCSK(η) = Z

(N)eCS(η) +

η3(1 − 2η)

(1 − η)3〈σ2〉

6〈σ3〉2×

(

〈σ2〉2 + 〈σ〉〈σ3〉)

, (15)

which does not coincide with Boublık’s extension to mix-tures of the CSK EOS [12]:

Z(N)BCSK(η) = Z

(N)BMCSL(η) +

η3(1 − 2η)

(1 − η)3〈σ2〉33〈σ3〉2 . (16)

Recently, Henderson and Chan (HC) have proposed amodification of the BMCSL EOS [13, 14] for the partic-

ular case of a binary mixture in which the concentrationof the large spheres is exceedingly small, starting from anasymmetric prescription for the radial distribution func-tions at contact. The resulting EOS, with σ1 ≥ σ2, is

Z(2)HC(η) = Z

(2)BMCSL(η) +

4ηx1

〈σ3〉

{

x1σ31

{

2(1 − η)2

(

1 − 〈σ2〉〈σ3〉σ1

)

+η2

2(1 − η)3

[

1 −( 〈σ2〉〈σ3〉σ1

)2]

+ exp

[

2(1 − η)2

( 〈σ2〉〈σ3〉σ1 − 1

)]

− 1

}

+η2x2

4(1 − η)3

( 〈σ2〉〈σ3〉σ2

)2

(σ1 − σ2)

×[

(σ1 + σ2)2 − ησ2(σ

21 + σ2

2 + σ1σ2)〈σ2〉〈σ3〉

]}

.

(17)

We shall now perform a comparison with the (veryfew) available computer simulation data. We begin withhard-disc mixtures. In figure 1 we display the packing-fraction dependence of the compressibility factor Z forthe SPT, W, eH and eSHY EOS, together with the sim-ulation results of Barrat et al. [7], for the binary mixturedefined by x1 = 0.351 and σ2/σ1 = 0.8. In this case,the performance of the eSHY EOS is outstanding andclearly superior to all the other choices. To complete thepicture, in figure 2 we present the results for the ratio ofthe fifth virial coefficient to the fourth power of the (ex-act) second virial coefficient as a function of the largerdisc concentration and for two size ratios. Here, the bestagreement with the numerical data of Wheatley [15] isobtained with the eH EOS, which is not very surpris-ing since in the one-component case (x1 = 1) it gives avery good estimate of this ratio. Nevertheless, the over-all trends including the position of the maximum are stillcaptured in all approximations.

As far as hard-sphere mixtures are concerned, thefollowing comments can be made. To our knowledge,only simulation results for binary mixtures have beenreported. The most recent data [16] indicate that the

0.48 0.50 0.52 0.54 0.56 0.583.5

4.0

4.5

5.0

5.5

6.0

η

Simulation eH eSHY SPT W

Z

FIG. 1: Compressibility factor as a function of the packingfraction for a binary mixture of hard discs defined by x1 =0.351 and σ2/σ1 = 0.8.

0.0 0.2 0.4 0.6 0.8 1.00.3

0.4

0.5

0.6

0.7

0.8

4

Numerical eH eSHY SPT

σ2/σ

1=0.2

σ2/σ

1=0.1

x1

B5/B

2

FIG. 2: Fifth virial coefficient as a function of x1 for twobinary mixtures of hard discs defined by σ2/σ1 = 0.1 andσ2/σ1 = 0.2.

BMCSL EOS underestimates the pressure as obtainedthrough simulation. In fact, the BCSK EOS is geared tocorrect this deficiency, at least for η < 0.5, in a similarfashion as the CSK EOS corrects the CS EOS. As a gen-eral trend, for η < 0.5, our extended EOS, namely theeCS and the eCSK, also go in the correct direction. More-over, in this density range, ZBMCSL < ZeCS < ZeCSK andZBMCSL < ZBCSK < ZeCSK. Although limited in scope,the results shown in figure 3 illustrate these features.Here we have considered an equimolar binary mixturewith size ratio σ2/σ1 = 0.6. As the differences betweenthe values predicted by the various EOS for the com-pressibility factor Z are very small, we have chosen topresent the results, including the simulation data of Yauet al. [14] (open circles) and Barosova et al. [16] (filled

Page 4: Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

4

0.1 0.2 0.3 0.4 0.5-0.02

0.00

0.02

0.04

0.06

0.08É

η

Z-Z

BM

CS

L Simulation eCSK eCS BCSK HC BMCSL

FIG. 3: Compressibility factor as a function of the packingfraction, relative to the BMCSL value, for an equimolar bi-nary mixture of hard spheres with σ2/σ1 = 0.6.

0.0 0.2 0.4 0.6 0.8 1.0

0.2

0.4

0.6

0.8

1.0

4

Numerical eCS BMCSL BCSK

σ2/σ

1=0.2

σ2/σ

1=0.05

x1

B5/B

2

FIG. 4: Fifth virial coefficient as a function of x1 for twobinary mixtures of hard spheres defined by σ2/σ1 = 0.05 andσ2/σ1 = 0.2.

circles), in terms of the packing fraction dependence ofZ − ZBMCSL. Despite the scatter of the simulation re-sults of Yau et al. [14], it is apparent that, dependingon the range, both the eCSK and eCS EOS seem to doa better job than either the BMCSL, BCSK or HC EOS(although in all fairness we should add that the equimolarcondition is beyond the scope for which the latter EOSwas originally devised). In fact, if one considers a higherdensity point computed by Yau et al. [14] (η = 0.59,Zsimul − ZBMCSL = 0.26) that is off the scale, it appearsthat the overall trend is better captured by the eCS EOS,although for η < 0.5 the eCSK should probably be thepreferred EOS. The results for the composition depen-dence of the fifth virial coefficient for a binary mixtureand two values of σ2/σ1 displayed in figure 4 also indi-cate that all the approximations lead to very good val-ues as compared to the recent numerical data of Encisoet al. [17], with a slight superiority of the BCSK EOSfor the region around the maximum. The HC and theeCSK results have not been included, since they are al-most identical to the ones of the BMCSL and the eCS,respectively.

In conclusion, it is fair to state that we have introduceda very simple and general recipe that allows one to geta reasonably accurate approximation to the EOS of amulticomponent mixture of d-dimensional hard-spheresfrom any reasonable EOS of the one-component system.It also seems that, as exemplified in the case of binarythree-dimensional hard-sphere mixtures, the more accu-rate the EOS of the one-component system, the betterresults the approximation yields.

Two of us (A. S.) and (S. B. Y.) would like to acknowl-edge partial support from the DGES (Spain) throughGrant No. PB97-1501 and from the Junta de Ex-tremadura (Fondo Social Europeo) through Grant No.PRI97C1041. M. L. H. wants to thank the hospitalityof Universidad de Extremadura, where the draft of thepaper was prepared.

[1] Carnahan, N. F. and Starling, K. E., 1969, J. chem.

Phys. 51, 635.[2] Boublık, T., 1970, J. chem. Phys. 53, 471; Mansoori,

G. A., Carnahan, N. F., Starling, K. F., and Le-

land, T. W., 1971, J. chem. Phys. 54, 1523.[3] Henderson, D., 1975, Molec. Phys. 30, 971.[4] Santos, A., Lopez de Haro, M., and Yuste, S. B.,

1995, J. chem. Phys. 103, 4622. For a didactic presen-tation, see also Lopez de Haro, M., Santos, A., and

Yuste, S. B., 1998, Eur. J. Phys. 19, 281.[5] Lebowitz, J. L., 1964, Phys. Rev. 133, 895.[6] Hansen, J.-P. and McDonald, I. R., 1986, Theory of

Simple Liquids (London: Academic Press).[7] Barrat, J.-L., Xu, H., Hansen, J.-P., and Baus, M.,

1988, J. Phys. C 21, 3165.[8] Reiss, H., Frisch, H. L., and Lebowitz, J. L., 1959,

J. chem. Phys. 31, 369; Helfand, E., Frisch, H. L.,

and Lebowitz, J. L., 1961, J. chem. Phys. 34, 1037.[9] Mulero, A., Cuadros, F., and Galan, C., 1997, J.

chem. Phys. 107, 6887.[10] Wheatley, R.J., 1998, Molec. Phys. 93, 965.[11] Kolafa, J., 1986, unpublished results. This equation

first appeared as equation (4.46) in the review paper byBoublık; T. and Nezbeda, I., 1986, Coll. Czech. Chem.

Commun. 51, 2301. We are grateful to Prof. Kolafa forproviding this information in a private communication.

[12] Boublık, T., 1986, Molec. Phys. 59, 371.[13] Henderson, D., Malijevsky, A., Labık, S., and

Chan, K.-Y., 1996, Molec. Phys. 87, 273; Yau, D. H.

L., Chan, K.-Y., and Henderson, D., 1997, Molec.

Phys. 91, 1137; Henderson, D., Soko lowski, S., and

Wasan, D., 1998, Molec. Phys. 93, 295.

Page 5: Equation of State of a Multicomponent D-dimensional Hard-sphere Fluid

5

[14] Yau, D. H. L., Chan, K.-Y., and Henderson, D.,1996, Molec. Phys. 88, 1237.

[15] Wheatley, R. J., 1998, Molec. Phys. 93, 675.[16] Barosova, M., Malijevsky, A., Labık, S., and

Smith, W. R., 1996, Molec. Phys. 87, 423.[17] Enciso, E., Almarza, N. G., Gonzalez, M. A., and

Bermejo, F. J., 1998, Phys. Rev. E 57, 4486.