Physica A 264 (1999) 15–39
Generalized hydrodynamics of binary uids
Kunimasa Miyazakia;∗ , Kazuo Kitaharab
a Department
of Physical Chemistry, National Institute of Materials and Chemical Research,
Tsukuba, Ibaraki 305-8565, Japan
b Department of Applied Physics, Tokyo Institute of Technology, Meguro-ku, Tokyo 152-8550, Japan
Received 12 May 1998
Abstract
The k-dependent hydrodynamic uctuations for binary classical uids are discussed. The aim
of this paper is to give the microscopic basis for the phenomenological formulation of hydrodynamics of binary uids which was explored by the present author. The generalized Langevin
equation formalism is used to derive closed equations for the uctuations of temperature, mass,
relative velocity, etc. Explicit expressions for the k-dependent thermodynamic susceptibilities are
also derived. Using these quantities, it is shown that the equations for uctuations reduce to the
macroscopic equations in the long wavelength limit. c 1999 Elsevier Science B.V. All rights
reserved.
PACS: 05.70.Ln
Keywords: Hydrodynamic uctuations; Binary uids
1. Introduction
In the previous paper (hereafter to be referred to as paper I) [1], we have generalized
conventional formalism of nonequilibrium thermodynamics of multicomponent uids to
the shorter time scale where inertial e ect of the di usional ow of each component
becomes important. There, the di usional ow is introduced as a new local extensive
variable by de ning the internal energy density as the part of energy, which does not
include the kinetic energy of macroscopic ows. The balance equation for an extensive
variable ai (t) has a general form as follows:
X
@S
@S
@ai X
Lij
+
;
(1.1)
{ai ; aj }
=
@t
@aj
@aj
j
j
Corresponding author. Present address: Theoretical Studies Division, Institute for Molecular Science,
Myoudaiji, Okazaki 444-8585, Japan. Fax: +81-564-53-4660; e-mail:
[email protected].
∗
0378-4371/99/$ – see front matter c 1999 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 8 - 4 3 7 1 ( 9 8 ) 0 0 4 4 6 - 4
16
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
where @S=@ai is the intensive parameter conjugate to ai and Lij is the generalized
Onsager coecient which has an operator form. Here, we have made use of discretized
variables rather than continuous eld variables for simpli cation. The rst term of
the right-hand side is the reversible part and the second is the irreversible part. The
coecient matrix {ai ; aj } has to be antisymmetric, namely {ai ; aj } = −{aj ; ai }, so that
the reversible terms do not contribute to the entropy production of the whole system.
The balance equation for the di usional ow which has not been known except for
a dilute case was thus deduced from the known equations for the total energy, mass
density, and momentum density by using the antisymmetric nature of {ai ; aj }.
However, there still remains some arbitrariness in the division of balance equations into the reversible part and irreversible part. The Onsager coecients satisfy
the well-known reciprocal relation Lij = i j Lji , where i = 1 if ai is the even function under time reversal transformation and i = −1 otherwise. More precisely, if the
Onsager coecients depend on a parameter B such as the external magnetic eld or
the velocity eld which are odd in the time reversion, we have Lij (B) = i j Lji (−B).
If one variable is even and another is odd in the time reversion and Lij does not
depend on B, the Onsager coecient becomes antisymmetric, i.e., Lij = −Lji . The
coecient thus has the same symmetry as {ai ; aj } and the term with this symmetry
does not contribute to the entropy production, i.e.,
X
@S @S
@S @S
dS X
=
Lij
=
;
(1.2)
Lij
dt
@a
@a
@a
i
j
i @aj
i; j
i; j; i j =1
P
where i; j; i j =1 denotes the summation over i; j with the same time reversal symmetry,
i.e., i j = 1. This implies that one may divide the reversible part and irreversible
part arbitrarily for the contribution from the variables with the di erent time reversal
symmetry. This arbitrariness may be avoided by going beyond phenomenological level
of description and resorting to microscopic description. According to the idea of linear
response theory, the Onsager coecients are related to thermal uctuations of ows
in the equilibrium state via the uctuation–dissipation theorem. One expects that this
theorem will avoid such arbitrariness in the division of reversible part and irreversible
part in the macroscopic evolution laws.
The primary purpose of this paper is to give microscopic basis of the hydrodynamic
equations derived in paper I which was based on a purely phenomenological level
of description. For this purpose, we shall consider the uctuations of thermodynamic
variables around equilibrium state. Wavelength-dependence of the uctuations is fully
taken into account. We shall focus here on binary classical uids which consist of
two types of atoms or molecules. The closed equations for the time-dependent correlation functions of the microscopic uctuations of the mass concentration ratio, mass
density, temperature, barycentric velocity, and relative velocity of each component are
derived by using the generalized Langevin equation formalism [2,3]. The equations
thus derived are expressed in terms of the k-dependent static correlation functions
and memory kernels. Another purpose of this paper is, though supplementary, to give
the microscopic expressions for k-dependent thermodynamic susceptibilities for binary
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
17
uids. It will be shown that the static correlation functions appearing in the generalized
Langevin equations are expressed in terms of k-dependent thermodynamic susceptibilities.
We shall show that, in the hydrodynamic limit, i.e., in the long wavelength and long
time limit, the generalized Langevin equations have the identical form as those derived
from hydrodynamical equations derived in paper I. The arbitrariness in phenomenological equations discussed above is, thus, clari ed by this procedure.
Such a procedure was carried out for simple one-component uids [3,4]. Extrapolation of the hydrodynamic description into the short wavelength or k-dependent regime
was referred to as the generalized hydrodynamics. The analysis in this paper is much
along the line of these works for one-component uids.
We shall also give formal expressions for the Onsager coecients. It is generally
known that the generalized Langevin formalism gives the Onsager coecients in terms
of the time correlation functions of the random forces whose dynamics are characterized by the “modi ed” Hamiltonian and therefore they are di erent from those which
are given by the Green–Kubo formulae. The latter give the Onsager coecients in
terms of the same type of time correlation functions of random forces whose dynamics are characterized by the total Hamiltonian. For one-component uids, the Onsager coecients derived from the generalized Langevin formalism reduce to those of
the Green–Kubo integrands in the hydrodynamic limit since dynamic variables chosen
are all conserved quantities (see for example Refs. [3]). In our problem, however, a
non-conserved variable, the relative velocity, is incorporated as one of thermodynamic
variables. This leads to discrepancy between correlation functions of the di erent formalisms.
This paper is organized as follows. In the remainder of this section, we summarize
the results of paper I. In Section 2, we introduce the proper microscopic expressions
for the uctuations of the internal energy density, mass density, momentum, and the
relative velocity and give their equations of motion. The k-dependent uctuations of
temperature, pressure, and chemical potential are introduced here. This generalization
is implemented such that they reduce to the macroscopic thermodynamic relations in
the small-k limit. Using these generalized thermodynamic quantities, k-dependent thermodynamic susceptibilities such as the compressibility or speci c heat are derived in
terms of the static correlation functions of uctuations of the thermodynamic variables.
In Section 3, the equations for the thermodynamic variables are derived by using the
generalized Langevin formalism. Explicit expressions for coecient matrices for the
reversible parts and irreversible parts of the balance equations are shown. Section 4 is
devoted to conclusions.
The content of this paper was brie y reported in the Second Tohwa University
International Meeting on Statistical Physics [5].
Let us summarize here results given in paper I [1]. We have shown that the equation
for the set of local extensive variables, the total energy density e, mass density for the
lth component l (l = 1; 2), total momentum, C ≡ 1 C1 + 2 C2 , where Cl is the velocity
of the lth component, and di usional momentum for component 1, J1 ≡ 1 (C1 − C),
18
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
are given by
1
@
2
(e) = −∇ · eC + P · C + + (1 − 2c)! J1 + Jq + C · ;
@t
2
(1.3)
@
l + ∇ · (l Cl ) = 0;
@t
(1.4)
(l = 1; 2) ;
@
(C) + ∇ · (CC) = −∇ · P − ∇ · ;
@t
(1.5)
@
J1 + C · ∇J1 + J1 ∇ · C + J1 · ∇C + (1 − c)∇ · {(1 − c)!J1 } − c∇ · {c!J 1 }
@t
1
@J1
+ ∇
+
;
(1.6)
=c(1 − c)T ∇ −
T
T
@t irr
where c ≡ 1 = is the mass concentration ratio for component 1 and ! = C1 − C2 is the
relative velocity between the two components. In the above expressions,
P
≡ p
+ c(1 − c)! !
is the generalized hydrostatic pressure, and = 1 − 2 is the chemical potential difference between two components. appearing in Eqs. (1.3) and (1.6) is an arbitrary
quantity which is not speci ed within the phenomenological level of description. Jq ; ;
and (@J1 =@t)irr are the irreversible parts of equations which are referred to as the
irreversible heat ow, the viscous tensor, and the friction force, respectively. Note that
the equation for another di usional momentum J2 is redundant since J1 = −J2 by
its de nition. In the framework of linear thermodynamics, the irreversible parts are
expressed as the linear combination of thermodynamic forces such as the temperature
gradient, the velocity gradient, and the relative velocity. We may consider the following
phenomenological expressions for the irreversible part as a typical example:
@v
@v
2
+
− ∇ · C − ∇ · C ;
= −
@x
@x
3
Jq = L00 ∇
@J1
@t
!
1
+ L01 −
;
T
T
!
1
;
+ L11 −
= L10 ∇
T
T
irr
(1.7)
where and are the shear viscosity and the bulk viscosity, respectively. L00 is the
coecient which is related to heat conduction and L11 is the friction coecient. L10 and
L01 are the cross coecients which are related to thermal di usion. Onsager reciprocal
relation and time-reversion symmetry yields L10 = −L01 . In view of Eq. (1.3), one
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
19
might expect that the heat ow carried by the mass di usion of constituents is given
by
2
X
{hl + 21 (Cl − C)2 }Jl = {h1 − h2 + 12 (1 − 2c)! 2 }J1 ;
(1.8)
l=1
where hl (l = 1; 2) is the molar enthalpy and, thus, should be the di erence of the
molar enthalpy for each component h1 − h2 as it was suggested by several authors [6,7]
in the framework of conventional nonequilibrium thermodynamics. It is found that it
is not the case when the faster time scales are concerned and the di usional ow is
taken as a thermodynamic variable rather than an irreversible ow. This point will be
discussed in Sections 3 and 4.
For later use, we shall give the linearized equations of Eqs. (1.3) –(1.6) around
equilibrium state. It is convenient to choose the mass concentration ratio, total mass
density, temperature, longitudinal parts of barycentric velocity and relative velocity as
independent variables and introduce the Laplace and Fourier transformed variables by
Z ∞ Z
dt
dr e−zt eik·r A(r; t) :
A(k; z) =
0
Then, we have a set of equations in the following form:
(z − i
+ ) · a(k; z) = ã(k) ;
(1.9)
where a(k; z) = (c; ; T; k̂ · C; k̂ · !) with k̂ ≡ k=k. ã(k) = a(k; t = 0) is the
initial value of the variable. We shall not consider the transverse parts of C and !
since they are independent of the dynamics of other variables. i · a stands for the
reversible part of the equations whose coecient matrix is given by
0
0
0
0
i c!
0
0
0
i v
0
(1.10)
i ≡ 0
0
0
i Tv i T!
;
i vc i v i vT
0
0
0
0
i !c i ! i !T
where each element is given by
i
c!
= ikc(1 − c) ;
i
v
= ik ;
T
H
; i T! = ikc(1 − c) ;
cv T
cv
1
@
; i v = ik 2 ; i v T = ik
;
i vc = ik
@ T; c
T
T
@
H
@
; i ! = ik
; i !T = ik ;
i !c = ik
@c T;
@ T; c
T
i
Tv
= ik
(1.11)
20
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
where cv = (@u=@T ); c is the speci c heat at constant volume, = −−1 (@=@T )p; c is
the thermal expansion coecient, T = −1 (@=@p)T; c is the isothermal compressibility
and
@
2
+ :
(1.12)
H ≡T
@T T ; c
The last term in the left-hand side of Eq. (1.9) is the irreversible part and its coecient
matrix is given in terms of Onsager coecients by
0 0 0
0
0
0 0 0
0
0
(1.13)
≡ 0 0 TT 0
T!
0 0 0
0
vv
0 0 !T 0
!!
with
TT
=
vv
=
!T
where
l
L00 2
k ;
cv T 2
lk
2
= −ik
T!
= ik
L01
;
cv T
;
L10
2
T c(1 −
c)
;
!!
=
L11
;
Tc(1 − c)
(1.14)
= (4=3 + )= is the longitudinal viscosity.
2. Thermodynamic derivatives
In this section, we introduce microscopic uctuations of extensive variables around
the equilibrium state. The uctuations of intensive parameters such as temperature,
hydrostatic pressure are also properly introduced. Thermodynamic susceptibilities are
shown to be expressed in terms of the proper combination of static correlation functions
of these uctuating variables.
2.1. De nition of the microscopic quantities and their equations of motion
We consider a system with volume V consisting of N1 and N2 particles of each
component. We assume that an intermolecular potential of the system is given by a
sum of pair potentials which depend only on the intermolecular distances:
U (r1(1) ; : : : ; r1(2) ; : : :) =
Nl X
Nm
2
1 X X
lm (|ri(l) − rj(m) |) ;
2
(2.1.1)
l; m=1 i=1 j=1
where ri(l) denotes the position of the ith particle of the lth component. The microscopic
de nition of mass density, momentum density of each component, and internal energy
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
21
density are de ned by
l (r; t) = ml
Nl
X
(r − ri(l) );
(l = 1; 2) ;
i=1
Nl
X
ui(l) (r − ri(l) ); (l = 1; 2) ;
(2.1.2)
i=1
Nl
2 Nm
2 X
X
1 XX
1
ml ui(l)2 +
lm (|ri(l) − rj(m) |) (r − ri(l) ) ;
(u)(r; t) =
2
2
(l Cl )(r; t) = ml
m=1 j=1
l=1 i=1
where ml and ui(l) = ṙi(l) are the mass and velocity of a particle of the lth component,
respectively. The averages over the equilibrium ensemble are given by
hl (r; t)i = l = ml nl ;
hCl (r; t)i = 0 ;
hu(r; t)i =
Z
2
1 X
3
nkB T +
nl nm dr glm (r)lm (r) ;
2
2
(2.1.3)
l; m=1
where h· · ·i denotes the grand canonical ensemble average over the phase space
(ri(l) ; ui(l) ). Here nl = Nl =V (l = 1; 2) is the number density of the lth component per
unit volume, and n ≡ n1 + n2 : glm (r) (l; m = 1; 2), which appears in the last equation,
is the radial pair distribution function de ned by
*N ; N
+
m
l
1 X
(l)
(m)
′
(r − ri + rj ) ; (l = 1; 2) ;
(2.1.4)
nl nm glm (r) =
V
i; j=1
where the prime on the sum indicates that terms with i=j; l=m are excluded. Hereafter,
we shall consider the uctuations around the equilibrium values in the wavevector
space, i.e., ak = ak − hai(2)3 (k). From the above de nition, we obtain the equations
of motion which the uctuations obey. The total mass uctuation ≡ 1 + 2 is
governed by the continuity equation.
@k
= ik · Ck ;
(2.1.5)
@t
where Ck ≡ cC1k + (1 − c)C2k is the uctuation of the barycentric ow given by
(
)
N1
N2
X
X
1
(1) ik·ri(1)
(2) ik·ri(2)
m1
ui e
+ m2
ui e
:
Ck =
i=1
i=1
The quantities without subscript k denote the average values, e.g., = hi. For the
mass concentration ratio ck , using ck = −1 {(1 − c)1k − c2k }, we have
@ck
= c(1 − c)ik · !k ;
@t
where !k = C1k − C2k is the uctuation of the relative velocity given by
!k =
N1
N2
(1)
(2)
1 X
1 X
ui(1) eik·ri −
ui(2) eik·ri :
n1
n2
i=1
i=1
(2.1.6)
22
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
For the barycentric velocity, the equation of motion is given by
@Ck
= ik · k
@t
with the total pressure tensor;
(lm) (lm)
Nl
Nm
2 X
2 X
X
X
r
r
1
ij
′ ij
Pk(lm) (rij(lm) ) eik · ri(l) ;
ml ui(l) ui(l) −
k=
(lm)2
2
rij
l=1 i=1
(2.1.7)
(2.1.8)
m=1 j=1
where the Greek indices are the Cartesian coordinates x; y; z. In this expression, we
have introduced
dlm (r) 1 − e−ik·r
:
Pk(lm) (r) ≡ r
dr
ik · r
The equation for !k is given by
@!k
= ik · k − Fk ;
@t
(2.1.9)
where
(11) (11)
N1
N1
X
r
r
(1)
1
1 X
ij
ij
(11) (11)
′
P
(r
)
eik·ri
ui(1) ui(1) −
k=
ij
k
(11)2
n1 i
2m1
rij
j=1
(22) (22)
N2
N2
X
r
r
(2)
1 X
1
ij
ij
(22) (22)
′
−
P
(r
)
eik·ri
ui(2) ui(2) −
ij
k
(22)2
n2 i
2m2
rij
(2.1.10)
j=1
and
F k=
N1 X
N1 X
N2
N2
rij(12) d12 (rij(12) ) ik·r(1)
rij(12) d12 (rij(12) ) ik·r(2)
1 X
1 X
i
e
e i :
+
1
2
rij(12) drij(12)
rij(12) drij(12)
i=1 j=1
i=1 j=1
(2.1.11)
For the internal energy density, since (u) = u + u, we have
1
1
uk = {(u)k − uk } = {(u)k − uk } ;
which can be written as
Nl
Nm
2 X
2 X
X
X
(l)
1
1
1
ml ui(l)2 +
lm (rij(lm) ) − ml u eik·ri :
uk =
2
2
l=1 i=1
(2.1.12)
m=1 j=1
Then, the time derivative of this is given by
@uk
= ik · qk ;
@t
(2.1.13)
where qk = qk(1) + qk(2) with
Nl
Nm
2 X
2 X
1
X
X
(l)
1
ml ui(l)2 +
lm (rij(lm) ) − ml u eik·ri
ui(l)
qk(1) =
2
2
l=1 i=1
m=1 j=1
(2.1.14)
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
23
Nl X
Nm
2
r (lm) rij(lm) (lm) (lm) ik·r(l)
1 X X
(l)
(m) ij
≡−
(ui + uj ) (lm)2 Pk (rij )e i
4
rij
(2.1.15)
and
q(2)
k
l; m=1 i=1 j=1
is the energy ow.
2.2. Temperature, pressure, and chemical potential
We shall here generalize uctuations of temperature, hydrostatic pressure, and chemical potential to the k-dependent regime. These uctuations are given in terms of linear
combination of the internal energy, mass density, and mass concentration ratio. The
coecients in the linear combination are determined such that they reduce to the usual
thermodynamic relation in the long wavelength limit. Using these variables, microscopic expression for thermodynamic derivatives (thermodynamic susceptibilities) are
derived.
The uctuation of temperature can be naturally introduced by using the fact that,
in the long wavelength limit, it is statistically independent of both the mass density
and mass concentration ratio (see Appendix A) [4,8]. We may write the k-dependent
temperature uctuation as
Tk = a1 (k){uk + a2 (k)k + a3 (k)ck } :
(2.2.1)
Coecients a2 (k) and a3 (k) are determined such that hTk −k i = hTk c−k i = 0.
Therefore, we have
a2 (k) = −
1
{(uk ; k )(|ck |2 ) − (uk ; ck )(k ; ck )} ;
(k)
a3 (k) = −
1
{(uk ; ck )(|k |2 ) − (uk ; k )(k ; ck )}
(k)
(2.2.2)
with
(k) ≡ (|k |2 )(|ck |2 ) − |(k ; ck )|2 :
(2.2.3)
Here we have de ned (ak ; bk ) ≡ V −1 hak b−k i and (|ak |2 ) ≡ (ak ; ak ).
On the other hand, a1 (k) can be determined by comparing Eq. (2.2.1) with thermodynamic relation which is valid for the long wavelength limit;
1 @u
1 @u
1
−
c :
(2.2.4)
T = u −
cv
cv @ T; c
cv @c T;
Therefore, it is natural to introduce the k-dependent speci c heat at constant volume
by a1 (k) ≡ 1=cv (k). The k-dependent speci c heat can be determined by using the
formula for uctuations (see Appendix A).
lim (|Tk | 2 ) = hT 2 i =
k→0
kB T 2
;
cv
(2.2.5)
24
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
where hT 2 i is the amplitude of the local temperature uctuation in the long wavelength
limit de ned by hT (r)T (r′ )i ≡ hT 2 i(r − r′ ). Using Eqs. (2.2.1) and (2.2.5), we
have
(|uk + a2 (k)k + a3 (k)ck |2 )
cv (k) =
kB T 2
=
{(|uk |2 ) + a2 (k)(uk ; k ) + a3 (k)(uk ; ck )} :
kB T 2
(2.2.6)
By comparing Eq. (2.2.1) with Eq. (2.2.4), the following relations have to be satis ed.
@u
@u
1
T
−p ;
lim a3 (k) = −
= 2
:
lim a2 (k) = −
k→0
k→0
@ T; c
T
@c T;
(2.2.7)
As for the uctuation of the hydrostatic pressure, we may determine it from the
de nition of the stress tensor which has been introduced in the derivation of the microscopic equation for the barycentric velocity in Eq. (2.1.8). We can separate the
stress tensor into a “Thermodynamic part” and a “Viscous part” as [4]
′
;
lk = pk + lk
(2.2.8)
where lk = k̂·· k̂ is the longitudinal component of the stress tensor. pk is determined
′
is orthogonal to T; , and c. We may write it as
such that the remainder lk
pk = b1 (k)Tk + b2 (k)k + b3 (k)ck :
(2.2.9)
Then, each coecient is given by
b1 (k) =
(lk ; Tk )
;
(|Tk |2 )
b2 (k) =
kB T (|ck2 |)
1
{(lk ; k )(|ck2 |) − (lk ; ck )(k ; ck )} =
;
(k)
(k)
b3 (k) =
1
kB T (ck ; k )
{(lk ; ck )(|k |2 ) − (lk ; k )(ck ; k )} = −
;
(k)
(k)
(2.2.10)
where use has been made of identities:
(lk ; k ) = kB T
and
(lk ; ck ) = 0 :
In the long wavelength limit, by comparing with the thermodynamic relation, the coecients bi (k) are to reduce to
1
@p
@p
= ;
=
lim b2 (k) =
;
lim b1 (k) =
k→0
k→0
@T ; c T
@ T; c T
lim b3 (k) =
k→0
@p
@c
T;
:
(2.2.11)
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
25
Using the relations exhibited above, one may derive the general expressions of thermodynamic derivatives of arbitrary quantities in terms of the static correlation functions
of microscopic quantities. If one takes the temperature, mass density, and concentration ratio as independent variables, uctuations of a quantity A are written in the long
wavelength limit as
X @A
ai :
A =
@ai
a =T; ; c
i
Each coecient can be derived by taking the ensemble average after multiplying a =
T; ; c from the left-hand side. Thus, we obtain
(Ak ; Tk )
@A
:
(2.2.12)
= lim
@T ; c k→0 (|Tk |2 )
This formula may be generalized as
(Ak ; Tk )
@A
:
= lim
k→0
@B ; c
(Bk ; Tk )
For other two coecients, we have
1
@A
{(Ak ; k )(|ck |2 ) − (Ak ; ck )(k ; ck )} ;
= lim
k→0 (k)
@ T; c
1
@A
{(Ak ; ck )(|k |2 ) − (Ak ; k )(k ; ck )}
= lim
k→0 (k)
@c T;
or more generally,
(Ak ; k )(|ck |2 ) − (Ak ; ck )(k ; ck )
@A
;
= lim
k→0 (Bk ; k )(|ck |2 ) − (Bk ; ck )(k ; ck )
@B
T; c
(Ak ; ck )(|k |2 ) − (Ak ; k )(k ; ck )
@A
:
= lim
k→0 (Bk ; ck )(|k |2 ) − (Bk ; k )(k ; ck )
@B T;
(2.2.13)
(2.2.14)
(2.2.15)
Applying the above argument for the entropy, pressure, and concentration ratio, the
adiabatic derivative is obtained;
(Ak ; pk )
(Ak ; lk )
@A
= lim
:
(2.2.16)
= lim
@B s; c k→0 (Bk ; pk ) k→0 (Bk ; lk )
Likewise, the isothermal and isobaric derivative is given by choosing the temperature,
pressure, and mass concentration ratio as independent variables,
(Ak ; ck )
@A
:
(2.2.17)
= lim
@B T; k→0 (Bk ; ck )
If we take A = u and B = in Eq. (2.2.16), then we obtain the virial expression of
the hydrostatic pressure;
Z
2
1 X
@u
dlm (r)
glm (r) :
= nkB T −
(2.2.18)
nl nm dr r
p = 2
@ s; c
6
dr
l; m=1
26
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
On the other hand, Eq. (2.2.17) with A = u; −1 and the enthalpy per unit mass, h ≡
u + p−1 , etc. and B = c leads to the molar energy, molar speci c volume, and molar
enthalpy, etc., respectively. The di erence between molar quantities per unit mass of
each component is de ned by [6]
@A
:
(2.2.19)
A1 − A2 =
@c T;p
Combining this equation with the relation; A = cA1 + (1 − c)A2 , one has
@A
@A
;
A2 = A − c
:
A1 = A + (1 − c)
@c T;p
@c T;p
(2.2.20)
For example, the partial speci c volume is given by putting A = −1 ,
1
(k ; ck )
(1 − c) lim
k→0 (|ck |2 )
2
R
1 + n2 dr{g22 (r) − g12 (r)}
R
:
=
m1 [n + n1 n2 dr{g11 (r) − 2g12 (r) + g22 (r)}]
v1 = −1 −
(2.2.21)
Speci c volume for component 2 is obtained by exchanging the indexes 1 with 2.
Following the similar way, we may also derive the microscopic expressions for
the thermal expansion coecients = −−1 (@=@T )p; c , the isothermal compressibility
T = −1 (@=@p)T; c and we shall nd that they are in accordance with Eqs. (2.2.2)
and (2.2.11).
Finally, let us consider the uctuation of the chemical potential. Instead of , we
shall consider =T for convenience;
= c1 (k)Tk + c2 (k)k + c3 (k)ck :
(2.2.22)
T k
ci (k) are determined such that Eq. (2.2.22) reduces in the long wavelength limit to
@
1 @
1 @
=
T +
+
c
T
@T T ; c
T @ T;
T @c T; c
1
=−
T
@u
@c
1
T + 2
T
T;
@p
@c
1
+
T
T;
@
@c
c :
T; c
Comparing this relation with Eq. (2.2.4), we have
c1 (k) =
1
a3 (k);
T
c2 (k) =
1
b3 (k) :
2 T
(2.2.23)
c3 (k) is derived from the relation in the long wavelength limit (see Appendix A);
@
kB T h2 i
;
(2.2.24)
=
@c T;
where ≡ (k = 0). Therefore, one has
c3 (k) =
kB (|k |2 )
:
(k)
(2.2.25)
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
27
The relations derived in this subsection will be used in the next section to see the
relevance of the microscopic equations for the uctuations to the phenomenological
equations.
3. Generalized Langevin equation formalism
In this section, we shall derive a set of closed equations which describes the dynamics
of uctuations of variables ak = (ck ; k ; Tk ; vk ; !k ), where vk = k̂ · Ck and
!k = k̂ · !k are the longitudinal part of the velocity elds. For this purpose, we shall
introduce the projection operator P de ned by [3]
PA(t) = (A(t); aik )(aik ; ajk )−1 ajk :
(3.1)
Following the conventional procedure which has been applied for one-component uids,
one obtains the equation for the correlation function matrix, in Laplace transformed
representation,
Cij (k; z) = (aik (z); ajk ) ;
which is given by
[z − i (k) + (k; z)]il Clj (k; z) = C̃ ij (k) ;
(3.2)
where C̃ ij (k) ≡ Cij (k; t = 0) is the initial value of the correlation function.
i
ij (k)
−1
≡ (ȧik ; alk )C̃ lj (k)
(3.3)
is the coecient of the reversible part of Eq. (3.2) and it is expressed in terms of
static correlation functions. On the other hand, (k; z) is the memory kernel which
describes the irreversible processes and is given by
ij (k; z)
= (U ′ (z)fik ; flk )(alk ; ajk )−1 ;
(i; j = T; v; !) ;
(3.4)
where
fik ≡ Qȧik = ȧik − i
ij (k)ajk
(3.5)
are the random forces and Q ≡ 1 − P is the projection operator which is orthogonal
to P.
U ′ (z) ≡
1
z − QiLQ
(3.6)
is the time-evolution operator de ned by the modi ed Liouville operator iQLQ. It can
be shown that one may add an arbitrary linear combination of aik to the random
forces in the kernels [9]. The simplest choice is, of course, +i ij (k)ajk so that the
random forces are given simply by fik = ȧik .
28
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Explicit expression of i (k) is given by
0
0
0
0
0
0
i
i (k) =
0
0
0
i
i (k) i (k) i
v
vT (k)
vc
i
!c (k)
i
! (k)
i
!T (k)
where
i
c! (k)
= ikc(1 − c) ;
i
v (k)
= ik ;
i
Tv (k)
=
0
i
c! (k)
Tv (k) i T! (k) ;
0
0
v (k)
0
0
0
T
(Ṫ k ; vk )
= ik
b1 (k) ;
(|vk |2 )
cv (k)
c(1 − c)
(Ṫ k ; !k )
= ik
[a3 (k) + h† (k)] ;
(|!k |2 )
cv (k)
ik kB T
1
(k ; ck ) = ik b3 (k) ;
i vc (k) = −
(k)
1
ik kB T
(|ck |2 ) = ik b2 (k) ;
i v (k) =
(k)
i
i
T! (k)
vT (k)
(3.7)
=
=i
Tv (k)
(3.8)
1
(|vk |2 )
= ikb1 (k) ;
(|Tk |2 )
ik kB T
(|k |2 ) = ikTc3 (k) ;
(k)
1
ik kB T
(ck ; k ) = ikTc2 (k) = 2 ikb3 (k) ;
i ! (k) = −
(k)
i
!c (k)
=
i
!T (k)
=i
T! (k)
1
(|!k |2 )
= ik [a3 (k) + h† (k)] ;
(|Tk |2 )
T
where use has been made of
(ċk ; !k ) = (!˙ k ; ck ) = ik
kB T
and
(˙k ; vk ) = (v̇k ; k ) = ik kB T :
Derivation of i T! (k) and i !T (k) is elucidated in Appendix B. In the above equations,
we have de ned h† (k) = h†1 (k) − h†2 (k) with the modi ed molar enthalpy
h†l (k) = ul† (k) +
1 †
p (k);
l l
(l = 1; 2)
of the lth component. Here, ul† (k) is the modi ed molar energy de ned by
)
(
Z
2
1 3
1X
†
nl kB T +
nl nm drlm (r)glm (r) ; (l = 1; 2)
ul (k) =
l 2
2
m=1
(3.9)
(3.10)
29
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
and pl† (k) is the partial pressure de ned by
pl† (k)
= nl kB T −
2
X
m=1
nl nm
Z
dr(k̂ · r̂)2 Re[Pk(lm) (r)]glm (r);
(l = 1; 2) ;
(3.11)
where r̂ = r=r. Note that in the long wavelength limit, sum over two components of
these quantities satis es
u = cu1† (k = 0) + (1 − c)u2† (k = 0);
p = cp1† (k = 0) + (1 − c)p2† (k = 0) ;
(3.12)
where u is the average of the internal energy density given by the last equation
of Eq. (2.1.3) and p is the virial expression of the hydrostatic pressure given by
Eq. (2.2.18). Therefore, we have h = u + p−1 = ch†1 (k = 0) + (1 − c)h†2 (k = 0).
Here it should be emphasized that the modi ed molar enthalpy is di erent from the
molar enthalpy de ned in the equilibrium thermodynamics, though both lead to the
same value if summed up over components. According to Eq. (2.2.19), conventional
de nition of the molar enthalpy is given by [6]
@h
@h
;
h2 = h − c
:
(3.13)
h1 = h + (1 − c)
@c T;p
@c T;p
Making use of Eq. (2.2.17), the microscopic expression for Eq. (3.13) is given by
(hk ; ck )
;
k→0 (|ck |2 )
(hk ; ck )
;
k→0 (|ck |2 )
h1 = h + (1 − c) lim
h2 = h − c lim
(3.14)
where hk = uk − −2 pk + −1 pk is the uctuation of the enthalpy density. After
laborious manipulation, one arrives at h1 = u1 + pv1 , where v1 is the molar speci c
volume given by Eq. (2.2.21) and u1 is the molar energy given by
u1 = u + (1 − c) lim
k→0
=
1
3 kB T
+
2 m1
2m1 A
+n2
2
X
(uk ; ck )
(|ck |2 )
" 2 Z
X
n
drnl 1l (r)g1l (r)
l=1
(−1)k+1 nl nm
k; l; m=1
Z
dr
Z
dr′ lm (r){glmk (r; r′ ) − glm (r)gk2 (r ′ )} ;
(3.15)
where
A ≡ n + n1 n2
Z
dr{g11 (r) − 2g12 (r) + g22 (r)} :
The expression for component 2 is given by exchanging subindices 1 and 2 in the
above expression. The derivation of Eq. (3.15) is elucidated in Appendix D. In the
30
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
above expression, glmk (r; r′ ) is the three-particle distribution function de ned by
*N N N
+
k
l
m
X
X
1 X
(l)
(m)
(l)
(k)
′
′
′
′
′
(r − ri + rj )(r − ri + rh ) ;
nl nm nk glmk (r; r ) =
V
h=1
i=1
j=1
(k; l; m = 1; 2):
(3.16)
It is obvious that h1 di ers from Eq. (3.9). Both quantities are identical only in the
dilute limit. In this limit, one may neglect correlations between particles and obtains
ul† = ul = 3kB T=2ml and h†l = hl = 5kB T=2ml .
Using the results in the previous section, it is easy to check that the matrix i (k)
reduces to i in Eq. (1.10) in the long wavelength limit if we put = limk→0 h† (k) =
limk→0 {h†1 (k) − h†2 (k)}. Thus, we may conclude that a proper heat ow due to the
di usional ow is given in terms of the modi ed molar enthalpy not of the molar
enthalpy in a usual sense.
On the other hand, the dissipative term has the following structure:
0 0
0
0
0
0
0
0
0 0
:
(3.17)
(k; z) =
(k;
z)
(k;
z)
(k;
z)
0
0
TT
Tv
T!
0 0
vT (k; z)
vv (k; z)
v! (k; z)
0 0
!T (k; z)
!v (k; z)
!! (k; z)
Comparing this equation with Eq. (1.14), we obtain the expressions for the Onsager
coecients
Z ∞
1 4
1
∗
+ =
lim
dthU ′ (t)lk · lk
i;
T 3
3kB V k→0 0
1
lim
3kB V k→0
Z
∞
1
lim
L01 =
3kB V k→0
Z
∞
1
lim
3kB V k→0
Z
∞
L00 =
L11 =
dthU ′ (t)qk · qk∗ i ;
0
(3.18)
′
dthU (t)qk ·
∗
J˙ 1k i
= −L10 ;
0
∗
dthU ′ (t)J˙ 1k · J˙ 1k i ;
0
where U ′ (t) = exp[iQLQt] is the modi ed time evolution operator and J1k = c(1 − c)
!k is the di usional momentum of the component 1. One may also obtain similar
expressions for LTv , Lv! but it is generally known that these coecients are negligible
for simple uids 1 [10].
1
As discussed in paper I, Lv! becomes important in such systems like colloidal suspensions, liquid 4 He
under -point.
31
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Note that the transport coecients Eq. (3.4) are di erent from the Green–Kubo
integrands which have the form of
ij (k; z) ≡
(U (z)ȧik ; ȧjk )
;
(|ajk |2 )
(i; j = T; v; !) ;
(3.19)
where U (z) = (z − iL)−1 is the Liouville operator. Both (k; z) and (k; z) are related
to each other by [3]
−1
1
· {(k; z) − i (k)}
(3.20)
(k; z) − i (k) = 1 − {(k; z) − i (k)}
z
or equivalently,
−1
1
:
(k; z) − i (k) = { (k; z) − i (k)} · 1 + { (k; z) − i (k)}
z
(3.21)
lim lim (k; z) = lim lim (k; z) :
(3.22)
The proof of these identities is shown in Appendix C.
For one component systems, these coecients become identical in the hydrodynamic
limit, i.e.,
z→0 k→0
z→0 k→0
This is because all thermodynamic variables ak = (k ; Tk ; vk ) are conserved variables and therefore ȧk is proportional to k. Thus, ; are found to be ∼ O(k 2 ) for
small-k limit and these quantities in the denominators in Eqs. (3.20) and (3.21) are
neglected in this limit.
For two component systems, however, since !k is not a conserved variable, the
above argument for equivalence between Green–Kubo integrands and memory kernels
does not hold.
In order to see the di erence, let us expand the inverse matrix in Eq. (3.20) up to
the second order in k.
−1
1 ′
1 − (k; z)
z
(
"
−1
−1
−1
1 ′
1
1
1
1
1
′
1 − 0
1 − ′0
· 1 +
· ′2 + 2 1 − ′0
= 1+
z
z
z
z
z
z
·′1
1
· 1 − ′0
z
−1
·
′1
)#
where ′ ≡ − i and ′i is the ith
obtain
−1
1 ′
′
′
· 0 + 1 −
= 1 − 0
z
1
+
z
(
1
1 − ′0
z
−1
·
′1
1
1 − ′0
z
−1
+ O(k 3 ) ;
(3.23)
order term in k of ′ . After some algebra, we
−1
)2
−1
1 ′
· 1 − 0
z
as k → 0 :
1 ′
z 0
−1
·
′1
1
· 1 − ′0
z
+ ′2
(3.24)
32
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Each coecient of the above expression is given by
cv
1 cv
1
|T! − i T! |2 −
|i Tv |2 ;
z − !! c(1 − c)T
zT
1
(T! − i T! )!v ;
Tv = Tv +
z − !!
z
!!
i T! ;
T! −
T! =
z − !!
z
z
v! ; as k → 0 ;
v! =
z − !!
c(1 − c)
1 cv
1
2
2
2
=
+
|
|
−
|i
;
|
+
|i
|
vv
vv
!v
Tv
v
z − !!
z T
3 T
cv
cv ∗
∗
;
;
!T = −
vT = −
T Tv
c(1 − c)T T!
z
1
∗
!! ;
;
!! =
!v =
c(1 − c) v!
z − !!
TT
= T T −
(3.25)
where ij∗ denotes the complex conjugate of ij . The denominator z − !! appearing
in the above expressions has a singularity at z = 0, since !! ∼ z as z → 0. This
singularity leads to a constant value for transport coecients
in the hydrodynamic
limit, while some of ij vanishes in this limit. The last terms of the equations for T T
and vv with singularities at z = 0 cancel out the similar contributions from T T and
vv , respectively.
4. Conclusions
In this paper, we have formulated the generalized hydrodynamics for binary
uids. Using the generalized Langevin equation formalism, the equations for the timedependent correlation functions of thermodynamic variables are expressed in terms of
k-dependent thermodynamic susceptibilities and memory kernels. In the hydrodynamic
limit, these equations becomes identical to the linearized equations which were derived from the phenomenological level of description. Arbitrariness in the paper I was
clari ed by analyzing microscopic uctuations and it was found that a free parameter
in Eqs. (1.3) and (1.6) is given in terms of the modi ed enthalpy limk→0 h† (k) =
limk→0 {h†1 (k) − h†2 (k)} de ned by Eq. (3.9) contrary to the conventional choice of the
usual molar enthalpy h = h1 − h2 . On the phenomenological level, cannot be determined. Only if we go into the microscopic level of description, we can nd that is
actually h† .
However, if one concerns longer time scales in which the inertial e ect of di usional
ow is neglected, we may choose arbitrarily. As shown below in detail, corresponding
to each choice of , we can de ne the proper random forces whose time correlation
functions give the transport coecients. In such long time limit, the di usional ow
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
33
plays the role of one of the irreversible currents rather than a thermodynamic variable.
As was pointed out by Bearman et al. [7], it becomes arbitrary to divide the irreversible
energy ow into the heat ow and the heat carried by the di usion from the statistical
mechanical standpoint. To see this, let us consider the linear relation between the
irreversible ow and thermodynamic forces. Neglecting the inertial terms in Eq. (1.6)
and using Eqs. (1.3) and (1.7), one obtains
1
+ L̃01 ∇ −
;
J̃ q = L̃00 ∇
T
T
1
+ L̃11 ∇ −
;
(4.1)
J1 = L̃10 ∇
T
T
where L̃ij are Onsager coecients which are used in the conventional nonequilibrium
thermodynamics and J̃ q ≡ Jq + h† J1 is the total irreversible ow of energy density.
The heat conductivity, thermal di usion ratio, and the di usion coecient are given in
terms of these coecients by
!
2
L̃11 @
@c
@c
L̃
L̃01
1
T;p
;
−h ; D=
= 2 L̃00 − 10 ; kT =
T
@ T;p L̃11
T
L̃11
(4.2)
respectively [6]. Here use has been made of a thermodynamic relation:
@h
@
2
:
= −T
h = h1 − h2 =
@c T;p
@T T
p; c
L̃ij are related to Lij by
L̃00 = L00 +
1
{L10 + c(1 − c)Th† }2 ;
L11
L̃01 = L̃10 =
c(1 − c)T
{L10 + c(1 − c)Th† };
L11
L̃11 =
{c(1 − c)T }2
:
L11
L̃ij can be expressed in the form of Green–Kubo formula as
Z ∞
1
lim
dthU (t)qk · qk∗ i ;
L̃00 =
3kB V k→0 0
1
lim
3kB V k→0
Z
∞
1
lim
L̃11 =
3kB V k→0
Z
∞
L̃01 =
∗
dthU (t)qk · J1k
i;
(4.3)
0
∗
dthU (t)J1k · J1k
i;
0
where U (t) = exp[iLt] is the time evolution operator rather than the modi ed operator
U ′ (t), since both the energy density and concentration ratio are the conserved quantities.
34
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Eq. (4.1) is the simplest way to de ne the heat ow. There is possibility of de ning
the heat ow in di erent ways keeping the Green–Kubo formulae Eq. (4.3) invariant
under this rede nition. If one divides the heat ow as
′
J̃ q = J̃ q + ′ J1 ;
(4.4)
where ′ is an arbitrary constant, then, the linear relation Eq. (4.1) is rewritten as
1
1
′
′
′
′
+ ∇
+ L̃01 ∇ −
;
J̃ q = L̃00 ∇
T
T
T
1
1
′
+ L̃11 ∇ −
+ ′ ∇
;
(4.5)
J1 = L̃10 ∇
T
T
T
′
where L̃ij is the Green–Kubo integral given by replacing the microscopic heat ow qk
in Eq. (4.3) by
qk′ = qk − ′ J1k :
(4.6)
A convenient choice of the parameter is ′ = h1 − h2 . Then, Eq. (4.5) becomes
1 ′
1
′
′
− L̃01 ∇T ;
J̃ q = L̃00 ∇
T
T
(4.7)
1 ′
1
′
− L̃11 ∇T ;
J1 = L̃10 ∇
T
T
where ∇T indicates the di erentiation keeping the temperature constant. This is the
choice adopted in most of the literatures [6,7].
Such arbitrariness does not appear when the friction force rather than the di usional ow is considered as an irreversible part of balance equations (1.3)–(1.6).
In the generalized Langevin equation formalism, incorporation of the di usional ow
as a thermodynamic variable corresponds to going down to the second level in the
hierarchy of the continued fraction structure of the equation [2] for the concentration
ratio.
In this paper, any explicit analysis of the transport coecients was not shown.
As was discussed in paper I, the formalism given here can a ord explaining fast
processes [11] or phenomena where di usion couples to convection in uids [12].
Acknowledgements
The authors are grateful to Prof. D. Bedeaux for illuminating discussions and useful advice. This work was nancially supported by the National Institute Postdoctoral
Fellowship from the Research Development Corporation of Japan (JRDC), Grant-in-Aid
for Scienti c Research (05245103), and Grant-in-Aid for Scienti c Research on
Priority Area (07240106) from Ministry of Education, Science and Culture, Japanese
Government.
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
35
Appendix A
According to the thermodynamic theory of uctuations [8], the probability density
for the uctuations of thermodynamic variables is proportional to
Z
dr{T (r)s(r) − p(r)v(r) + (r)c(r)} ;
(A.1)
exp −
2kB T
where v ≡ −1 is the speci c volume per unit mass. If we choose (T; ; c) as
independent variables, we have the following quadratic form:
Z
dr{T (r)s(r) − p(r)v(r) + (r)c(r)}
Z
dr
+2
=
@
@
1
cv 2
T (r) + 3 2 (r)
T
T
(r)c(r) +
T; c
@
@c
T;
2
c (r)
)
:
(A.2)
This implies that the uctuations are completely local and its static correlation functions
have a general form as
hai (r)aj (r′ )i = hai aj i(r − r′ ) :
(A.3)
The amplitudes of the variances of all of these variables are given by
kB T @
kB T
kB T 2
2
2
;
; h i =
; hc2 i = 4
hT i =
cv
@c T;
T
kB T
hci = −
@
@
;
@
@c
hTi = 0;
hTci = 0
(A.4)
T; c
where
1
= 3
T
@
@c
T;
−
2
:
T; c
Or equivalently, we may have
1
kB T hc2 i
kB T 2
;
;
=
2
3
hT i
T
@
@
kB T hci
kB T h2 i
;
=−
=
@ T;
@c T;
cv =
with
≡ h2 ihc2 i − hci2 :
As can be seen from Eq. (2.2.3), is the long wavelength limit of (k).
(A.5)
36
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Appendix B
T! (k)
In this appendix, derivation of i
Eq. (2.2.1), we have
and i
!T (k)
in Eq. (3.8) are given. Using
(!˙ k ; Tk ) = a1 (k){(!˙ k ; uk ) + a2 (k)(!˙ k ; k ) + a3 (k)(!˙ k ; ck )}
ik kB T
a1 (k)
(!˙ k ; uk ) + a3 (k) ;
=
i k kB T
(B.1)
where use has been made of (!˙ k ; k ) = 0 and (!˙ k ; ck ) = ik kB T=. The rst term
in the braces on the right-hand side is
(!˙ k ; uk ) = −
(!k ; u̇ k )
i k kB T
i k kB T
=−
1 ˆ ˆ
k k ((v1
kB T
(2)
− v2 k ); (q(1)
k + q k )) :
k
(B.2)
Each term is calculated as
"
2
c
kB T
m1
1
(1)
hv1 k q −k i =
×5
N1
V
Vn1
2
ml
kB T
+V
m1
kB T
=
1
"
(
2
1X
n1 nl
2
l=1
Z
dr1l (r)g1l (r) − 1 u
2
1X
5
n1 kB T +
n1 nl
2
2
l=1
Z
)#
#
dr1l (r)g1l (r) − 1 u
(B.3)
and
1
1
hv1 k q(2)
−k i = −
V
4Vn1
×
*
N1
X
rij(lm)2
(ui(l) + uj(m) )
l; m=1 i=1 j=1
i′ =1
rij(lm) rij(lm)
Nl X
Nm
2 X
X
(1)
ik·ri′
ui(1)
′ e
(lm) (lm) −ik·ri(l)
P−k
(rij )e
+
+
"
* (lm) (lm)
Nl X
Nm
N1 X
2 X
′
rij rij
1 X
(lm) (lm) ik·r (l′ l)
(1) (l)
hui′ ui i
P−k (rij )e i i
=−
4Vn1 ′
rij(lm)2
i =1 l; m=1 i=1 j=1
(m)
i
+ hui(1)
′ uj
*
rij(lm) rij(lm)
rij(lm)2
′
(lm) (lm) ik·r(l′ l)
P−k
(rij )e i i
+#
* (lm) (lm)
"
+
Nl X
Nm
N1 X
2 X
rij rij
kB T X
(lm) (lm) ik·r (1l)
′
P−k (rij )e i i
i′ i 1l
=−
4Vn1 ′
rij(lm)2
i =1 l; m=1 i=1 j=1
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
*
+ i′ j 1m
rij(lm) rij(lm)
rij(lm)2
Nm
N1 X
2
kB T X X
=−
4Vn1
l=1 i=1 j=1
+
=−
*
rij(1l) rij(1l)
rij(1l)2
"*
(lm) (lm) ik·r (1l)
P−k
(rij )e i′ i
rij(1l) rij(1l)
rij(1l)2
(1l) (1l) −ik·rji(1l)
P−k
(rij )e
37
+#
+
(1l) (1l)
P−k
(rij )
+#
Z
2
r r
kB T X
dr 2 Re[Pk(1l) (r)]g1l (r) :
n1 nm
2n1
r
(B.4)
l=1
−1
hv2 k q(2)
V −1 hv2 k q(1)
−k i and V
−k i can also be evaluated in the same way. Summing
up all of these terms, we arrive at
−
1 ˆ ˆ
k k ((v1
kB T
k
(2)
†
− v2 k ); (q(1)
k + q k )) = u1 (k) +
1 †
1
p (k) − u2† (k) − p2† (k)
1 1
2
(B.5)
with the modi ed molar internal energy de ned by
1
ul† (k) =
l
(
2
3
1X
nl kB T +
nl nm
2
2
Z
)
drlm (r)glm (r)
(B.6)
(k̂ · r)2
Re[Pk(lm) (r)]glm (r) :
r2
(B.7)
m=1
and the modi ed partial pressure
pl† (k) = nl kB T −
2
X
m=1
nl nm
Z
dr
This is the result shown in Eq. (3.9).
Appendix C
In this appendix, derivation of Eqs. (3.20) and(3.21) is shown. For the stationary
system, we have an identity:
ij (k; t) ≡ (ȧik (t); ȧlk )(alk ; ajk )−1 = (aik (t); ajk )(alk ; ajk )−1
= C ij (k; t)Clj−1 (k; t = 0) :
(C.1)
38
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
Laplace transforming both sides of this identity, we obtain
Z ∞
Z ∞
t) · C−1 (k; t = 0)
dt e−zt (k; t) = −
dt e−zt C(k;
(k; z) =
0
0
= [Ċ(k; t = 0) + zC(k; t = 0) − z 2 C(k; z)] · C−1 (k; t = 0)
= i (k) + z − z 2 C(k; z) · C−1 (k; t = 0) :
(C.2)
Comparing this with Eq. (3.2), we arrive at Eqs. (3.20) and (3.21).
Appendix D
In this appendix, derivation of Eq. (3.15) is shown. Using Eq. (2.2.20), one has
u1 − u2 = lim
k→0
= lim
k→0
(uk ; ck )
(|ck |2 )
1
{(1 − c)((u)k ; 1k ) − c((u)k ; 2k ) − u(k ; ck )} :
2 (|ck |2 )
(D.1)
We shall consider the rst term in the braces of the last line of the above equation.
Let us divide this into the contribution from the kinetic energy and from the potential
term as
((u)k ; 1k ) = ((u)k ; 1k )k + (( u)k ; 1k )p :
(D.2)
The kinetic term is written as
((u)k ; 1k )k =
Nl
2 X
X
ml
l=1 i=1
2
(l)
ui(l)2 eik·ri ; m1
ik·ri(l)
e
i=1
l=1
ik·ri(1)
′
e
i′ =1
Nl
X
2
3m1 kB T X
=
2
N1
X
;
N1
X
e
ik·ri(1)
′
i′ =1
!
!
Z
Z
31 kB T
ik·r
ik·r
1 + n1
dre g12 (r) :
dre g11 (r) + n2
=
2
As for the potential term, we have
N1
2 NX
l ; Nm
X
X
(1)
(l)
1
eik·ri′
lm (rij(lm) )eik·ri ; m1
(( u)k ; 1k )p =
2
′
l; m=1 i; j=1
2
m1 1 X
=
2 V
m=1
i =1
*N ;N
m
l
X
i; j=1
+
lm (rij(1m) )
(D.3)
K. Miyazaki, K. Kitahara / Physica A 264 (1999) 15–39
2
1 X
+
V
l; m=1
*N ;N N
m X
l
1
X
′
′
i; j=1
lm (rij(lm) )e
ik·rii(1l)
′
i′ =1
" 2
Z
1 X
nl
dr1l (r)g1l (r)
=
2
39
+
l=1
+
2
X
l; m=1
nl nm
Z
dr
Z
′
dr′ lm (r)eik·r glm1 (r; r′ ) :
(D.4)
(( u)k ; 2k ) can also be evaluated in the same manner. (|ck |2 ) and (k ; ck ) in
Eq. (D.1) was already derived in Eq. (2.2.21). Combining these results, one obtains
Eq. (3.15).
References
[1] K. Miyazaki, K. Kitahara, D. Bedeaux, Physica A 230 (1996) 600.
[2] H. Mori, Prog. Theor. Phys. 33 (1965) 423, ibid, 34 (1965) 399.
[3] J.P. Boon, S. Yip, Molecular Hydrodynamics, Dover, New York, 1991; J.-P. Hansen, I.R. McDonald,
Theory of Simple Liquids, Academic Press, London, 1986.
[4] P. Scho eld, Proc. Phys. Soc. 88 (1966) 149; P. Scho eld, in: H.N.V. Temperely, J.S. Rowlinson, G.S.
Rushbrooke (Eds.), Physics of Simple Liquids, North-Holland, Amsterdam, 1968, p. 563.
[5] K. Miyazaki, K. Kitahara, D. Bedeaux, in: M. Tokuyama (Ed.), Proc. 2nd Tohwa University Int. Meeting
on Statistical Physics, World Scienti c, Singapore, p. 98.
[6] S.R. de Groot, P. Mazur, Nonequilibrium Thermodynamics, Dover, New York, 1984.
[7] R.J. Bearman, J.G. Kirkwood, J. Chem. Phys. 28 (1958) 136.
[8] L.D. Landau, E.M. Lifshitz, Statistical Physics, Part 1, 3rd ed. Pergamon, New York, 1980.
[9] L.P. Kadano , P.C. Martin, Ann. Phys. 24 (1963) 419.
[10] I.M. de Schepper, E.G.D. Cohen, C. Bruin, J.C. van Rijs, W. Montfrooij, L.A. de Graaf, Phys. Rev. A
38 (1988) 271.
[11] K. Miyazaki, Phys. Lett. A 231 (1997) 409.
[12] M. Okamura, in: V. Krinsky (Ed.), Nonlinear Phenomena and Complex Systems, Kluwer Academic
Publishers, Boston, in press.