Transport Properties of a Dense Fluid of Molecules Interacting with a

Transport Properties of a Dense Fluid of Molecules Interacting with a Square-Well Potential. H. Ted Davis, and K. D. Luks. J. Phys. Chem. , 1965, 69 (...
0 downloads 0 Views 1MB Size
T R A K S P O PROPERTIES RT OF A DEXSEFLUID

869

Transport Properties of a Dense Fluid of Molecules Interacting with a Square-Well Potential

by H. Ted Davis and K. D. Luks Department of Chemical Engineering, University of Minnesota, Minneapolis, Minnesota (Receiaed August 28, 1964)

The theory developed by Davis, Rice, and Sengersl for the transport properties of a model fluid whose molecules interact according to a square-well potential is outlined herein and numerical computations based on the theory are presented. A perturbation technique is given for obtaining the equilibrium pair correlation function using the correlation function of hard spheres as the unperturbed term. Numerical calculations are performed for liquid argon leading to the conclusion that the square-well model furnishes a good first approximation to simple liquids. Also examined is a modified hard-sphere theory whereby the basic mechanism of transport is assumed to arise from hard-core collisions while the attractive interactions increase the effective collision cross section by increasing the magnitude of the pair correlation function. This modified hard-sphere theory is much better than the unmodified hard-sphere theory.

I. Introduction and Remarks I n recent years many investigators have attempted to develop a kinetic theory of liquids and dense gases which would have the same simplicity and degree of rigor enjoyed by the kinetic theory of dilute gases. The general approach to this problem has been to try to obtain, from the Liouville equation for the N-body distribution function, kinetic equations or equations of change for the singlet distribution function f(R1, p1; t ) and the doublet distribution function fi(R1, Rz, pl, p2; t ) , where R, and pi are, respectively, the position and momentum of the ith particle in an N-particle system. The singlet distribution function for particle i denotes the probability of finding a t time 1 the ith particle a t position R, with momentum p, regardless of the state of the other particles in the system, while the doublet distribution function denotes the probability of finding at time t two particles i and j a t positions R, and R, with momenta p, and p, regardless of the state of the other particles in the system. The most general kinetic theory of dense fluids to date based on the Liouville equation is the theory begun by Kirkwood2 and later extended by his coworkers. Kirkwood took the Liouville equation as a starting point and derived a Fokker-Planck type trans-

port, equation for liquids. He introduced a dissipative mechanism (which in the case of a dilute gas is the binary collision) by time-smoothing reduced distribution functions (e.g., the singlet and doublet distribution functions) over an interval of time T , where T is macroscopically short but microscopically long. That is, T must be long relative to the interval during which there is appreciable correlation between the total intermolecular force acting upon a specified molecule a t the beginning and end of that interval. Using analogies to the theory of Brownian motion, Kirkwood was able to derive a Fokker-Planck equation in which a friction constant { is to be identified with the phenomenological friction coefficient appearing in the Langevin equation. The friction constant is defined in terms of the intermolecular interactions averaged in time over T and over an equilibrium ensemble in phase space. Using the Fokker-Planck equation (for both the singlet and doublet distribution functions) all the transport properties of a system of molecules can be found in terms of { and the local thermodynamic variables of the system. Thus, in principle, the transport proper(1) H. T.Davis, S. A. Rice, and J. V. Sengers, J . Chtm. P h y s , 35, 2210 (1961), (2) J. G. Kirkwood, ibid., 14, 180 (1946).

V o l u m e 69, ‘V’umber 9 N a r c h 1966

H. TEDDAVISA N D K. D. LUKS

870

ties of a fluid are determined simply by evaluating {, a task not a t all easy for liquids, however. The biggest criticism to be levied against Kirkwood’s theory is that he had to assume that the momentum exchange. between particles is small during the time interval T ; that is, it was assunied that molecules moved through the fluid experiencing small random deflections. Clearly, this assumption excludes the possibility of core collisions (strong short-range repulsive interactions) between molecules resulting in strong deflections of the interacting molecules. Rice and Allriatt? have developed a theory which avoids the difficultly encountered by Kirkwood’s theory. They have assumed that a iiiolecule nioving through a dense fluid will undergo a motion in which it experiences a rigid-core collision followed by a “Brownian motion” during which it undergoes small random deflections due to the soft attractive interactions with its neighbors. Thus the iiiotion is determined by rigid-core collisions between which “Brownian motion” occurs. Rice and Allnatt subsequently obtained a modified Boltziiiann equation in which the hard-core collisions were treated as in the theory of the dense rigid-sphere fluid while the soft attractive interactions were handled in the Fokker-Planck approximation. Several nunieriral comparisons of theory and experiments 6 indicate that the Rice-Allnatt theory gives indeed a quantitative description of transport in simple liquids. However, in order to calculate the transport coefficientsfrom the Rice-Allnatt theory, one must have accurate values for the intermolecular potential and, niore importantly, one must know quantitatively the equilibrium correlation function as a function of the intermolecular distance at various temperatures and pressures. This latter demand is rather hard to meet since the pair correlation function has not been deterniined very accurately experimentally and the theoretical methods are still approximate and difficult. The practical difficulties involved in utilizing the more rigorous Rice-Allnatt theory seeiii to justify the exploitation of iiiodels which are cruder representations of reality but which can be treated more easily and exactly inathematically. Rigid spheres are the simplest molecules, but having no attractive interaction, they are missing one of the major features of a real liquid. Thus the next simplest choice is the “square-well” inodel fluid coinposed of particles which interact as rigid spheres at an interniolecular distance of say u1, have a constant attractive potential energy e for interinolecular distances between u1 and u2, and do not interac’t for interniolecular distances greater than uz.

Recently, Davis, Rice, and Sengersl derived a niodiThe Joiirnal of Physical Chemistry

fied Boltzmann equation for a ‘(square-well” fluid, from which they were able to formulate in analytic form the coefficient of viscosity and thermal conductivity. I n sections I1 and I11 we outline briefly their work, giving also the self-diffusion coefficient for a “square-well” fluid which was obtained by Longuet-Higgins and Valleau.’ In the remaining sections are presented numerical computations of the transport coefficients for liquid argon for a rather wide range of temperature and pressure (8T-150°1.VR, m f(Rl,pl;t)

=

J(f)

(2-2)

The quantities R,, pi, and m appearing in eq. 2-2 and 2-3a are, respectively, the position, momentum, and mass of the ith particle. g = [(pf/m) - (p2/7?Z)] Volume 69. Y u m b e r 3

March 1966

H. TEDDAVISAND K. D. LUKS

872

is the relative velocity of particles 1 and 2, k is a unit vector in the direction of their line of centers a t the instant of collision, and dk is the differential solid angle. In eq. 2-3a, the pair correlation function g(Rl,Rz,pl,pz;t) has been approximated by the local equilibrium pair correlation function g(R1,R2). Throughout eq. 2-3a, the position vector R2is evaluated a t the contact position for the particular partial collision considered, e.g. , g(Rl,R1 ulk) denotes the pair correlation function for particles 1 and 2 when particle 2 is found a t position alk relative a2k) is the pair to the position of particle 1. g(Rl,Rl correlation function at u2 6 and g*(R1,Ri uzk) is the pair correlation function a t u2 - 6 with 6 .-t O+. Under the assumption that g(R,,R,) is the local equilibrium pair correlation function, we have the relation g(R1,Rl uzk)/g*(R1,R1 a h ) = exp(-e/kT). For each of the four collision integrals of eq. 2-3a, one can calculate the nionientuni change from the binary dynaniics of the four different types of collision.

+

+

+

Type2.

=

g.k

>0

p(2rmkT)-"' exp[ -

-m(g.k)k

=

(2-3b)

I

p(Rl;t)u(R1;0 = S !!!fo(Rl,pl;t)dp, m

The scalar k appearing always in the combination kT is the Boltziiiann constant. 4 is the perturbation which in the linear approximation was shown to have the form

Type

A

< -(4e/m)"'

Ap1(3) = - 2 {g.k

+ i ( g . k ) Z- %}k m

B (2-3d)

+ B:OR,u

(2-7)

A and B were evaluated in terms of Sonine polynomials according to a variational method introduced by Hirschfelder, et al." To the first approximation A and B were found to be

m Type 3. g . k

mu)2 2mk T -

p(Ri it) = Sfo(Ri,Pi $)dPi

4 = A.VR, In T pi - pi'

(PI

I n this expression p , T , and u represent, respectively, the local values of the nuniber density, temperature, and hydrodynamic velocity. These quantities are defined in the usual nianner by

Typel. g . k > O Ap1'"

=

+

+

+

fo

=

= a(')(j/2

- Wz)W

bo("(WW - '/3W21)

(2-8)

where

4. - (4e/m)"' < g . k < 0 Apl(*) = -m(g.k)k

(2-3e)

The constraint g . k > 0 for types 1 and 2 indicate that the particles are initially approaching each other. The constraint g . k < -(4e/m)'" indicates that for a type 3 collision the particles are initially moving apart with a relative kinetic energy greater than e, while for a type 4 collision, - - 4 e / v ~ ) ~ '< * g . k < 0 indicates that the particles are moving apart and have a relative kinetic energy less than E. Finally, one should recall that Apl(') = - A ~ z " ' because of the conservation of nionientum. Equation 2-2 was linearized by a procedure following the Chapman-Enskog'O solution of the MaxwellBoltzniann equation. The singlet distribution function was written as

f

=

sou + cp)

(2-4)

where fa is the local equilibrium singlet distribution function defined by the relation T h e Journal of Physical Chemietry

with the abbreviations

(10) S. Chapman and T. G. Cowling, "The Mathematiral Theory of Nan-Uniform Gases," Cambridge University Press, New \-ark, N. T., 1953. (11) J. 0. Hirschfelder, C. F. Curtiss, and R. B. Bird, "Molecular Theory of Gases and Liquids," John Wiley and Sons, Inc., Xew York. N. T., 1954.

TRANSPORT PROPERTIES OF

A

DENSEFLUID

873

pk

+

= - P ~ T ~ O ( ' ) [ ~ / - Z ( V R , VB~U') U

-

1 / 3 V ~ l . U l ]- p k T l

(3-4)

where VR,U+ is the transpose of VR,u. Comparison of the phenomenological equations (2-13)

R

(2-14)

= az/ai

q p

= -[p

+

('/39K

+ KV)VE,T

= -(KK

+

'/37?V

-

+

@)vE1'u]l

~ ( V K VV)(VR,U

W is the reduced peculiar velocity and is given by

(3-5)

+

+ VB,U+)

(3-6)

leads to the results (2-15)

2k T

(3-7)

g(ul) and g ( a z ) are the equilibrium radial distribution functions for RpJ= c1 0 and a2 0, respectively.

+

+

111. Transport Coefficients of a Dense "Square-Well" Fluid Energy and momentum transfer can be accomplished by two molecular mechanisms in a fluid. I n the one case, kinetic energy and momentum may be transported by a single molecule from one point to another in a fluid simply by the molecular motion of the molecule. This mechanism gives rise to the so-called kinetic contribution to the heat flux and pressure tensor. In the other case, transport arises from collisional transfer whereby energy and inonientuni may be passed across an arbitrary surface by binary collisions involving no net transfer of molecules across the surface. This type of mechanism gives rise to the so-called intermolecular force contribution to the fluxes. Let us discuss first the kinetic contributions to the heat flux and pressure tensor denoted respectively as q k and pk. These quantities are simply the mean values of the molecular energy flux '/zm(pl/m - u)~. (pl/m - u) and the molecular momentum flux (pi mu)(pl/m - u ) , where pl/m - u is the velocity of particle 1 with respect t,o the hydrodynamic velocity u. In terms of the singlet distribution function the mean values of these fluxes are

The perturbation solution for f(R1,pI;t) given in section I1 is substituted into eq. 3-1 and 3-2 and the integrations are carried out leading to the expressions

(3-8) where KK and KV are, respectively, the kinetic and interniolecular contributions to the heat conductivity, TIC and rjv are the kinetic and intermolecular contributions to the shear viscosity, @ is the bulk viscosity, and P is the hydrostatic pressure. There is no kinetic contribution to the bulk viscosity @. Let us turn next to the intermolecular contributions to the fluxes. Chapman and Cowlinglo show that for rigid spheres the rate of transfer of momentum or energy due to the intermolecular forces is equal to

*

where is the molecular property transferred. I n the case of niomentuni transfer \k' = -Ap. The singlet distribution functions in eq. 3-9 are expanded in a Taylor's series about the point R1 and to the linear approximation eq. 3-9 beconies

In

Kifil -

k(g.k)dkdpldp2 (3-10)

I n eq. 3-10, terms of order VR~$Ihave been neglected and the abbreviation f(Rl,p,;t) = f(j) has been introduced. In the case of molecules interacting according to the square-well potential, the expression for the rate of transfer of momentum or energy has the same form as that of eq. 3-9 except that there are four types of collisions giving rise to four terms like the one in eq. 3-9. Thus the intermolecular force contribution to the pressure tensor is Volume 69. Number 3

March 1.966

H. TEDDAVISA N D K. D. LUKS

874

-

(3-17) with12

-

(3-18) -

(3-1 1) where the nionientuni changes Apl"' for each type of collision are given in eq. 2-3b-2-3e. Expanding the singlet distribution functions in eq. 3-1 1, keeping the linear approximation as indicated by eq. 3-10, and finally performing the indicated integrations leads to

where we have denoted by E the rate of strain tensor, i.e., E = VR,u V R , U + . If eq. 3-12 is combined with eq. 3-4 and the result is compared with the phenonienological equation, eq. 3-6, it is easily seen that

+

(3-13)

(3-14) (3-15) with (3-16) where we have inserted into eq. 3-14 the explicit value of bo'') as given by eq. 2-10. The quantities b, R, X and $were defined in section I1 (eq. 2-11 to 2-14). Proceeding with an equation for the interiiiolecular heat flux similar to eq. 3-11, it is found after some manipulation that the thermal conductivity is The Journal of Physical Chemistry

In eq. 3-14 and 3-17, the second term inside the braces is the same result obtained by Longuet-Higgins and Valleau for the zeroth approximation to the transport coefficients. The zeroth approximation is that for which the perturbation to the singlet distribution function is zero, ie., a") = bo(') = 0. The other contributions to q and K arise from sums of the kinetic ternis V K and KK and the interniolecular contribution due to the perturbation to the singlet distribution function (the quantity multiplying the factor e 1/3VB,.u1in eq. 3-12 is this sort of contribution). We shall not give herein a derivation of the self-diffusion coefficient for a "square-well" theory, but rather we shall simply take the results of Longuet-Higgins and Valleau,' namely

Equation 3-19 was obtained by assuming that the velocity correlation dies out exponentially with time. In the limit of a zero attractive potential (E = 0), eq. 3-19 reduces to the first approximation for dense hard spheres.

IV. Some Numerical Computations In order to calculate one of the transport properties froni the above equations, one must know the values of the parameters u1, m, E , g(ul), and g(u2) in addition to equation of state data. Values of ul, u2, and E can be determined empirically from gaseous data. In principle, g(ul) and g ( u 2 ) can be calculated from equilibrium statistical mechanics if one is given a, u z , and E. However, to the author's knowledge, there have been no theoretical calculations of the correlation function for a dense "square-well" fluid. Furthermore, there is to date no way of performing the calculations exactly. On the other hand, if one of the transport coefficients is known experimentally, then the predicted equation of state (eq. 3-13) and the predicted (12) I n their publication, Davis. Rice, and Sengersl suhmitted the expressions for x and r ) in much less transparent forms than presented in eq. 3-14 and 3-17. T h e simpler and more symmetrical versions presented herein are due to Dr. J. Sengers, who rearranged the original results.

TRASSPOHT PROPERTIES OF A DEXSEFLUID

875

equation for the known transport coefficient can be solved simultaneously for q(g1) and q(u2), presuming, of course, that g l , u2, E, and the PVT data are known for the fluid in question. This semiempirical procedure is the one we have followed in this section. I n section VI we shall present an approximation scheme for calculating directly the pair correlation function. Ikenberry and Rice5 and Sengers13 have recently made rat her extensive measurements of the thermal conductivity of' several siniple fluids. We have used the former's measurements to compute the correlation functions and from these we predicted the viscosity coefficient and diffusion coefficient for liquid argon over temperatures and pressures ranging from 87 to 150°K. and 1 to 500 atm. Values for ul, cz,and ~ / k were taken from virial coefficient data" to be, respectively, 3.16 8.,5.86 8.,and 69.4OK. Equation of state data were taken from tables compiled by for the liquid along the saturated vapor pressure curve, while the generalized charts of Hougen, Watson, and RagatzX4bwere used in computing the remainder of the P V T data used. The predictcld values for the viscosity and selfdiffusion coefficients are presented 'in Table I. The observed viscosity coefficients were taken from a table compiled by C00k.l~" The observed diffusion co-

Table I : Calculated and Observed Values of the Viscosity Coefficient I) and the Self-Diffusion Coefficient D for Liquid Argon at Various Temperatures T and Pressures P. Experimental Values for the Thermal Conductivity Were Vsed to Determine the Pair Correlation Functions

(calcd.),

v x 101 (obsd.),

poise

poise

7l

T,

F,

OK.

atrn

87 5 91 0 105 6 120 3 136 3 149 4 91 04 91 27 91 06 105 57 105 47 105 35 120 25 120 46 120 4 8 136 26 135 78 135 87 149 63 149 60

1 0

15 50 12 3 26 8 45 5 23 9 lOG 0 495 8 25 0 loci 3 495 1 24 7 100 1 500 8 27 9 loo 0 500 5 lOG 0 500 8

x

101

0 199 0 194 0 159 0 113 0 0927 0 0599 0 196 0 203 0 356 0 163 0 173 0 217 0 127 0 142 0 203 0 0942 0 122 0 176 0 109 0 143

0.246 0.220 0.148 0.113 0.077 0.051

D X 106 (calcd.), cm.'/eec.

1.70 1.81 2.54 3.46 4.96 6.57 1.80 1.76 0.955 2.45 2.36 1.76 3 33 3.15

D X 106 (obsd.), cm.(/sec.

2.27 2.50 4.20 6.03 8.26 10.00 2.32 1.72 4.13 3.07 5.96 4.96

2.32

5.09 4.01 3.00 4.33 3.68

8.35 6.37 8.16

0 EXPERIMENT

A MODIFIED HARDSPHERE CALC' S ,O

.a a C

/

/

0'4t' I

I

I

1

efficients were taken from a recent publication of Saghizadeh and RiceS6 As is illustrated in Figure 1, the agreement between the predicted and observed viscosities is quite good where we have data for comparison (the data presented in Figure 1 are for the liquid along the saturated vapor pressure curve). The agreement between observed and calculated self-diffusion coefficients is less satisfying than that for viscosities. The predicted values are too small by a factor of about 1.6 in most cases. However, as is shown in Figure 2, the correct temperature and pressure dependence is predicted. The data presented in Figure 2 are for values along the saturated vapor pressure curve. The dashed curve was calculated by Saghizadeh and Rice6 using experimental viscosity data to determine q ( r 1 ) and g(az) while the open triangular points were taken from Table I. The lack of agreement of the magnitude of the calculated diffusion coefficients perhaps reflects the in(13) J. V. Sengers, Thesis, University of Amsterdam, Amsterdam, Holland (14) (a) G. A. Cook, "Argon, Helium and the Rare Gases," Interscience Publishers, New Tork, N. Y., 1961; (b) 0. A. Hougen, K. M , Watson, and R. A. Ragata, "Chemical Process Principles," Vol. 2, John Wiley and Sons, Inc , New York, N. T , 1962.

Volume 69. Y u m h e r 3

M a r c h 1966

H. TEDDAVIS. 4 S D K. D. LUKS

876

10.0 DIFFUSION COEFFICIENT

VISCOSITY ISOBARS SQUARE-WELL CALC'S

EXPERIMENT

0-0-

.35

SQUARE-WELL CALC'S FROM VISCOSITY DATA A

\

6.0

\\

o\o-----\

.MI-

-_. 2

I

1.5

adequacies of' the model or perhaps arises froni the assumption that the autocorrelation function of momentum decays exponentially in time. I t would be interesting to calculate D ci la Chapman-Enskog from the transport equation (eq. 2-2). Actually, DRS estiiiiated the diffusion coefficient by expanding the collision integral into a Fokker-Planck type differential equation thus obtaining a diffusion coefficient. They improved on the Longuet-Higgins and T'alleau D by about 3%. However, there is no reason to make the approsiniations leading to a Fokker-Planck equation since D may be found by the same procedure employed to obtain 9 arid K . In view of the agreement of the predicted arid experiiiiental viscosities as indicated by Figure 1, one may conclude that the square-well model is a good first approxiniatiori to simple fluids.

100 atm.

-0-0-

-A-A-

SQUARE-WELL CALC'S FROM OBSERVED THERMAL COND.

1

,

1

\

MOOlFlED HARD SPHERE CALC'S

'. \ \\

--I--.-500 atm.

--e-.--

\-