Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

The Thermochemistry of a Reacting Mixture

of Elastic Materials with Diffusion


RAY M. BOWEN
Communicated by C.-C. WANG

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
2. Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3. Equations of Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4. The Entropy Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5. Constitutive Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6. Consequences of the Entropy Inequality . . . . . . . . . . . . . . . . . . . . . 112
7. Equilibrium Consequences of the Entropy Inequality . . . . . . . . . . . . . . . 118
8. Material Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

1. Introduction
The study of various thermomechanical theories of mixtures has received a
large amount of attention in recent years. Considerable attention has been given
to mixture theories where chemical reactions have been omitted. The work of
BOWEN [l], GREEN & NAGHDI [2], CROCHET & NAGHDI [3], GREEN & STEEL [4],
STEEL [5, 6] and MILLS [7] fall into this category. Mi3LLER [8] and INGRAM &
EmNGEN [9] have discussed certain aspects of chemically reacting fluids. BOWEN [10]
has presented a thermochemistry of viscoelastic materials of the Voigt type
where chemical reactions were present but diffusion was omitted. The arguments
presented here contain a thermochemistry of a reacting mixture of elastic mate-
rials with diffusion.
The various papers referenced above suffer from the unfortunate and confusing
feature that they often use conflicting entropy inequalities. This fact has been
discussed in some detail by BOWEN [1], MOLLER [8] and BOWEN & WIESE [11]
and will not be labored here. I simply want to point out that the balance equa-
tions and the entropy inequality used here are the same as the ones used by
BOWEN & WlESE [11], and they specialize to the ones used by BOWEN [10]. The
work of BOWEN & WIESE can be regarded as an improvement of the diffusion
theory presented by BOWEN [1]. The balance equations used here are essentially
the same as those originally proposed by TRtmSDELL [12] and TRUESDELL& TOU-
PIN [13, Sections 159, 215 and 243]. TRUESDELL'Soriginal formulation was later
generalized by KELLY [14] and ERINGEN & INGRAM [15]. The entropy inequality
used here (see Section 4) is motivated and discussed by BOWEN & WlESE [11].
It has also been discussed in a recent work by TRUESDELL [16, Chapter 5]. This
98 R.M. BOWEN:

inequality is not the same as the one proposed by GREEN & NAGHDI [2]. The
entropy inequality used by MULLER [8, Section 3] differs from the one used here
in that MLILLER leaves the entropy influx arbitrary and includes an additional
constitutive equation for this quantity. Of the various mixture theories currently
being examined in the literature, the theories of BOWEN [1, 10], MOLLER [8] and
BOWEN & WIESE [11] agree with the classical mixture theories. GREEN & NAGHDI
[17] have indicated that their theory must he generalized so as to enjoy this fea-
ture.
As stated above, this work concerns the formulation of a thermochemistry
of a mixture of reacting elastic materials with diffusion. The most interesting
general feature of this formulation is that when chemical reactions are present
one can no longer regard the density fields as determined by the deformation
fields. Thus, the constitutive assumptions must reflect this independence and, even
in elementary theories, allow for a dependence on a larger number of variables.
As a direct consequence of the independence of density and deformation fields,
it is no longer natural to regard the isotropy group of each constituent to be
contained in the unimodular group. It can only be said that the isotropy group
of each constituent is in the general linear group, i.e the group of nonsingular
linear transformations of a vector space onto itself. It turns out that this larger
symmetry is essential in order for the theory presented here to be specialized to
known results for mixtures of chemically reacting fluids with diffusion.
In Section 2, I outline the kinematics of motion appropriate for a mixture
of 9l bodies. In Section 3, I present the equations of balance of mass, momentum
and energy. Section 4 contains a discussion of the entropy inequality, and the
constitutive assumptions which define the mixture are stated in Section 5. Sec-
tions 6 and 7 contain the restrictions required by the entropy inequality. In Sec-
tion 8, the ideas regarding material symmetry mentioned above are formulated.

Notation
The direct notation is used throughout this paper. To the extent that it is
practical, I have followed the notation used by TRUESDELL & NOLL [18]. In
particular, points in the Euclidean 3-space g and vectors in its translation space
q / a r e denoted by Latin minuscules: x, q ..... Linear transformations from q/into
q / a r e indicated by boldface Latin majuscules: T, F, C,... and form a linear space
~ . If T is a linear transformation T r denotes its transpose, T-1 its inverse, tr T
its trace and det T its determinant. The notation or163 ~ denotes the elements of
s which are regular (invertible) and 6e denotes the elements of s which are
symmetric. The space of tensors on q/ of order p (p > 1) is denoted by ~ . For
simplicity, ~ and any other vector space of the same dimension are regarded as
the same. For example, q / a n d ~ and ~ and ~ are regarded as the same. The
set of tensors in ~ + 1 (p > 2) which satisfy the symmetry condition

G(u, vt, v2, ... , v p ) = G ( u , t~.(1), t~.(2) . . . . . v. (p)) (1.1)

for all u, r 1..... vp in q / a n d for all permutations a is a subspace of ~ + 1 which


is denoted by Sap+1. IfA is in ~ ( p > 2 ) and (i,j) are a pair of integers such that
Thermochemistry of a Reacting, Diffusing Mixture 99

l < i < j < p , the tensor CtjA in ~ - 2 defined by

(CijA)(u,, . . . , u i - 1 , ui+ l . . . . , u j - 1 , u j + l , . . . , Up)


a (1.2)
= ~ A(ul . . . . . ui-1, ek, Ui+l . . . . . llj_l, e~, uj+ l .... , up),
k=l

where (e 1, e a, ea) is a basis of W, is the concentration of A with respect to (i,j).


For example, if F is an element of La, then
C12F=trF. (1.3)
With the help of the contraction operator an inner product can be defined on
~ . If A and B are in ~ , then A | B is a tensor in Y2 p. The complete contraction
operator C is defined by
C(A | B ) = C12 ... CI.p-, ClpCl,v+l(A | B), (1.4)
p tinges

and it can be shown that C(A | B) defines an inner product on 5rp. The compo-
nent version of C(A | B) is
C(A | B) = Ak,k2...kp Bkl k2...kp. (1.5)

It can also be shown that


C(T| F) = tr(Tr F) (1.6)
for T, F in ~ =La. Two other contraction operators which we shall use are
defined by
(~(A | B | D ) = C24 C24(A | B | D) (1.7)
and
C(E | H)=Cla C14(E | H), (1.8)

where A, E and H are in ~33 and B and D are in ~22. In components, these opera-
tors can be defined by
C(A | B | D)= Ai jk BjS Dkmei | es | e=
and (1.9)
C ( E | H ) = E i j k H i j s e k | gs.

Finally if A is in ~ and B is in ~ + 1, then A[B] is a vector defined by

A In] = C12 ... Cl ~ Cl, ~+ ~(a | B). (1.10)


p tiptoes
In components, we can write
A [B] = Ak~k=...kp Bkl k~...k~q eq. (1.1 1)

The gradient and divergence with respect to spatial coordinates are denoted
by grad and div, respectively. The gradient with respect to material coordinates
is denoted by GRAD. A quantity corresponding to a particular material in the
mixture is identified by placing a labeling index directly below the symbol for
the quantity. The labeling indices for the constituents are always taken from
100 R.M. BOWEN:

the list: a, b, r b, e. Summations over the constituent indices are always indicated
by a summation sign. When there is no possibility of confusion, the index on
the summation sign shall be omitted.
As the reader will quickly learn, the notation used in mixture theories is
cumbersome. It is for this reason that I have been intentionally careless about
confusing functions and their values. In fact, unless some possibility of confusion
exists, I have repeatedly used the same symbol for a function and its value and,
in some cases, for several functions and their common value.

2. Kinematics
In this section the kinematics of motion for a mixture of 91 bodies ~ , a =
tl

1 ..... 9l is briefly discussed. Each body is taken to be a set endowed with the
structure prescribed by NOEL [19]. The usual kinematical concepts for theories
of single materials apply to each constituent. A configuration of the body ~Q is
a diffeomorphism Z
a
of ~a into 8. A motion of ~t~ is a one parameter family of
configurations, It, where t denotes the time parameter. If X denotes a particle
a ii

of the body ~ , its position at the time t is


x=zt(X)=z(X, t), (2.1)
tl a a Ct

where the function of Z, defined on ~ x ( - oo, oo), is defined by (2.1)2. A reference


tl fl

configuration for the body ~a is a fixed configuration r.t l The value of ra at X is


written
X=ra(X ) . (2.2)

It follows from (2.1) and (2.2) that


x = •a (Ira (X), t) = ~, (X, t), (2.3)
Q

where the deformation function X, is defined by (2.3)2. The inverse of I , exists


II n

for each t since/~ and r - 1 are diffeomorphisms. It shall be assumed here that
tl a

;t~ and its inverse ~-1 are of class C a.


a a

The velocity, acceleration and deformation gradient for the particle X at the
time t are defined as follows a
(x, t)
jg=--
a t3t '
a2Z (X, t) (2.4)
tl a
=TV-
and
F = GRAD Z,,(X, t).
~I t~ a
Thermochemistry of a Reacting, Diffusing Mixture 101

As in (2.4)1 and (2.4)2, throughout this paper we shall adopt the notation, intro-
duced by TRUESDV,LL [12], that a backward prime affixed to a quantity denotes
material differentiation following the motion of the body indicated by the sub-
script. It is easily shown that the material derivative of a function ~ along the
path of X a
is given by

~' _ ~ (z(X, t), t)_ OTt (x, t)+(gra d ~(x, t))~, (2.5)
=--ffi-- a Ot

where ~(x, t) is an element of ~ . As a result of the assumptions of smoothness


made about the functions Z~ and Z~-1, it follows that
a a

IdetFI
a
> 0. (2.6)

The velocity gradient for X


a
at the time t is

La = grad ~a (x, t). (2.7)

It i s a routine matter to show that (2.7), (2.4)1 and (2.4)3 imply that

L=FF-
a aa
a. (2.8)

If pa denotes the density of ~a at (x, t), then

P=Z~ (2.9)

is the density of the mixture at (x, t). The mass concentration at (x, t) is

a
c=plp,
a
(2.1o)
and, by (2.9),
Y'c=l. (2.11)

We shall regard the density p to be given by a non-negative C 2 function P~


such that a a
ap=P~(X, t). (2.12)

It is permissible for some of the densities to be zero at a given time so long as


they are all not zero at this time.
The velocity of the mixture i at (x, t) is defined by

= 1/p E g ~(x, t). (2.13)

The diffusion velocity ua for aX, which at t is at x, is defined by


o
U=X--X.
a a
(2.14)

8 A r c h . R a t i o n a l M e c h . A n a l . , Vol. 34
102 R.M. BOWEN:

It follows from (2.13) and (2.14) that


(2.15)

The velocity gradient for the mixture at (x, t) is


L= grad ~ (x, t). (2.16)
As (2.13) suggests, differentiation along the curve defined by (2.13) shall be
denoted by a superimposed dot. Also, if ~ is a tensor field of order p, then

~ = d~(x, t)+(gra d ~(x, t))x. (2.17)


at
It follows from (2.17), (2.5) and (2.14) that

~ - ~b=(grad ~U(x, t))u


(2.18)
= ( G R A D ~(X, t))( -a au ,
where ~(X, t)= ~(a/~(X, t), t).

Associated with each chemical element in nature is a positive real number


called its molecular weight. The molecular weight of a given constituent is ex-
pressible as a linear combination of the atomic weights of a finite number of
atomic elements. If we assume that the body ~ is formed from 9~ atomic elements,
then the molecular weight rn* is given by
9.I

rn*= E T~w', (2.19)

where w" is the atomic weight of the ~th atomic element and T~ is a nonnegative
integer that represents the mass of the ~th atomic element in the a th constituent
per unit mass of the ~th atomic element. Equivalently, T~ represents the number
of moles of the ~th atomic element in one mole of the o th constituent. The rank
of the matrix [T~] is denoted by ~ < min (9/, 9.0. By eliminating the dependent
columns of [T",], we can write (2.19) in the form

m"= Z S~ m', (2.20)


0t=l
where @ =rank [S~].
3. Equations of Balance
The equations of balance of mass, momentum and energy for a mixture of
9/ bodies have been discussed at length by TRUESDELL [12, 16], TRUESDELL &
TOUPIN [13], KELLY [14], ERINGEN & INGRAM [15], BOWEN & WIESE [11] and
GREEN & NAGHOI [20]. The equations presented here are essentially those for-
mulated by TRUESDELL. The equations of balance are the following:
a) Balance of mass for the a th constituent

(Pa IdetFI) = IdetFI ~, (3.1)


Thermochemistry of a Reacting, Diffusing Mixture 103

b) Balance of mass for the ~th atomic element

E S*."6/m" O, = (3.2)

c) Balance of linear momentum for the a th constituent

p ~ =div T+p b+~, (3.3)


r r r Ct

d) Balance of linear momentum for the mixture


~ ( ~a + ~aJr r (3.4)

e) Balance of moment of momentum for the a th constituent

a Q r
(3.5)

f) Balance of moment of momentum for the mixture

((x - o) A/~ + , ~ + ~(x -- o) A Jr = 0, (3.6)


a tl a G

g) Balance of energy for the a m constituent

Pa! = tr ( F - ' T r F ) - div q + p r + ~ (3.7)


tt Ct tl QQ

and
h) Balance of energy for the mixture
(3.8)

In equations (3.1) through (3.8), Q~ is the mass supply for the a t~ constituent,
T
a
is partial stress on the a th constituent, ba is the external body force density for
the a th constituent, ~ is the momentum supply or local body force on the a th con-
a

stituent, h?/is
Q
the moment of momentum supply or local body couple for the a th
constituent, e is the internal energy density for the a th constituent, q is the partial
cl (1

heat flux vector, Qr is the heat supply density for the a th constituent and ~tt is the
local energy supply for the a th constituent. So as to insure the proper invariance
of (3.7) under a change of observer, the energy supply ~Q is assumed to include
the rate of work of the body couples.
It is important to observe that if we multiply (3.2) by ~ " and sum the resulting
expression, then (2.20) shows that
E~=0. (3.9)

Equation (3.9) is the statement of balance of mass for the mixture. Several equiv-
alent versions of the remaining equations of balance can be obtained from (3.3)
*8
104 R.M. BOWEN:

through (3.8). For example, (3.9) and (2.14) can be used to replace (3.4) by

Z(~+~u)=O. (3.10)

Equation (3.3) can be used to replace (3.5) by


IQ= T - T r, (3.11)
a (1 a

and, if we use (3.4), (3.6) becomes


l (3.12)
Finally, (2.14), (3.9) and (3.10) can be used to write (3.8) in the form
E~=0, (3.13)
where
~=~+u./7+~(~+ 89 (3.14)
a a a a~a a a

The presence of a body couple for each constituent so as to yield (3.11) was
suggested, but not explored, by TRUESDELL & TOUPIN [13, Section 215]. This
feature was introduced formally by KELLY [14], and, later, by ERINGEN & ING~M
[15] and GlU~N & NAGHDI [20].
Equations (3.9), (3.10), (3.12) and (3.13) represent the equations of balance
for the mixture. The more familiar forms of these equations are the following:
a) Balance of Mass
Op
Ot +divp ~ = 0 , (3.15)

b) Balance of Linear Momentum


p ~--div T+p b, (3.16)
c) Balance of Moment of Momentum
T= T r (3.17)
and
d) Balance of Energy
p ~=tr TL-div q + pr + E p u . b, (3.18)

where T is the stress tensor for the mixture defined by

T= E (aT - Pa~ | ~)= E (aT- P, u | ~), (3.19)

b is the external body force density defined by


1
b=-~Ep, b , (3.20)
Thermochemistry of a Reacting, Diffusing Mixture 105

e is the internal energy density for the mixture defined by

1 / 1 ,2k 1 .2 1 ~ 1 2

r is the external heat supply density for the mixture defined by


1
r =-p- Z Para (3.22)

and q is the heat flux vector for the mixture defined by

q=r~r a(TTjc--q--Pa ~ a-
=E(aq_ TT U+ pa(~_1_89~2) au)" (3.23)

The equivalence of (3.9) and (3.15) is easily seen if we use (3.1), (2.9), (2.13),
(2.18)1, the identity
Idet Fa I = ]det F[
a
tr ~'F-
aa
1 (3.24)

and (2.8). Equation (3.17) follows from (3.12), (3.11) and the definition (3.19).
The equivalence of (3.16) and (3.10) and (3.18) and (3.13) follow by a straight-
forward but lengthy set of calculations. The interested reader can refer to TRUF_S-
DELL d~ TOUPIN [13, Sections 215 & 243] for the details. Following TRUESDELL..
the inner part of the internal energy density is defined by

1 ~ Paae' (3.25)

and the inner part of the stress is defined by

Tt = ~ T. (3.26)

From (3.17) and (3.19), we see that T~ is symmetric. In terms of the above quan-
tities, (3.18) can be written

P~=tr(TxL)-div(q-89 U) ~'au" (aaPX-pb)89-


act Eaa
Cu2"^ (3.27)

If we now use (3.3) and (3.26), (3.27) can be written

pez=tr~(F-1TrF)-divk-~u.~- 89 (3.28)
where
k=q-~Pa(-TT/pa+89 ~'(aq+PeUaaa)" (3.29)

It is possible to show that (3.2) is equivalent to

~1m~ ~, P~j~,, (3.30)


at= l
106 R.M. BOWEN:

where [P,~] is any matrix of rank ~R= 9 l - ~ such that

~ S~P,"=O # = 1, . . . , R (3.31)
, ~=1,..., S "

The integer ~R represents the maximum number of independent chemical


reactions taking place in the mixture. The matrix [P~] is called the s t o i c h i o m e t r i e
matrix, and the quantity j,, a = 1, ..., 91 is the reaction rate for the ath chemical
reaction. A discussion of the ideas leading to (3.30) and (3.31) can be found in
the paper by BOWEN [21]. With the exception of certain arguments in Section 7,
for the most part there is no need for us to use (3.30) directly.

4. The Entropy Inequality


In this section I shall postulate an entropy inequality for the mixture and then
examine the various forms it can take. Each constituent is assigned an entropy
density r/ and a temperature 0. The entropy density for the mixture is defined
to be a a
t/=lEpa~, (4.1)

and the temperature for the a th constituent is assumed to be given by a positive


valued C 2 function O~ such that
a

0a = O~
a
(X,
a
t). (4.2)

The entropy inequality is the postulate that


pr
p tl + div ~'.(q/0 +pa ~ u ) - ~ ~ > 0 (4.3)
a

at all points x and for all times t. The inequality (4.3) is the one discussed by
TRUESDELL [16] and BOWnN & Wmsn [11]. An equivalent, but possibly more
suggestive, version of (4.3) is

Opq
~ ' b~ d l v ( P"a a ~ xa ) §
.

a a
pr
aa

~ ]
) (4.4)

By using (3.7) and (3.14), we can write (4.4) in the form

[-Pa ( + ~ 0 ) + t r ( ~ - x T r Fa) - q /a0 "ag r a d 0 - u a ' p - ~a( ~ ba +a8 9 a a (4.5)


fl

where ~, is the partial Helmholtz free energy density defined by


a

~Q b =ae - 0a ,Q. (4.6)


Thermochemistry of a Reacting, Diffusing Mixture 107

In the special case when each constituent is constrained so that each consti-
tuent has the same temperature 0, we can use (3.29)2, and (4.6) to write (4.3)
in the form
p~+div[l(k-2P~ua)l-Pr/O>O. (4.7,

The equation that (4.5) specializes to is

_p(~,+t/b)+tr2F_ 1Tr~_h. grad0 ^ 1 ^ 2


a a 0 2 . u. . p - - -2 2 . acu - d i v 2 p ~. u > a0 , ( 4 . 8 )
where
h=k-2pO.
and (4.9)
1

The quantity ~k~ is the inner part of the free energy density. Equation (4.8) can
be derived either from (4.5) or, more directly, from (4.7), (3.28) and (4.9)2. We
are ultimately interested in examining a theory of reacting elastic materials in
which there is but one temperature for the mixture. The calculations are simpli-
fied if (4.8) is rewritten in the form

_~m_p~lO_trV2pF_,KF_h. grad0 2 a u . ~ - ~ - 1~ u 2 __>0, (4.10)


aa a a 0 a ~. aa

where K
a
is the chemical potential tensor defined by
Ka=~I-Tr/pa , (4.11)

and ~ is the free energy of the a th constituent per unit volume defined by
II

y=p~. (4.12)

The derivation of (4.10) from (4.8) also made use of (3.15), (2.18)1, (2.14) and
(4.9)2. Note, in passing, that (4.11), (4.9)2 and (3.26) imply that

l~pK=r T~/p. (4.13)


P
In an earlier work BOWEN [1] took the chemical potential tensors as primitive.
In effect, the partial stresses are taken as primitive here. If (4.11) is substituted
into (4.7) we obtain the entropy inequality postulated by BOWEN. An interesting
expression for the body couples ~I1/ i s obtained if we substitute (4.11) into (3.11).
The result is
~=Pa(K-Kr). (4.14)

As mentioned by BOWEN & WIESE [11, Section 3], equation (4.7) is not equiv-
alent to the entropy inequality postulated by GREEN &: NAGHDI. Their inequality
can be obtained if the t e r m 2 p ~ k u is omitted from (4.7), (4.8) and (4.9)1. Note,
gt a I1
108 R.M. BOWEN:

in passing, that the form of the energy equation (3.28) which, when combined
with (4.7), yields (4.10) is

pO~+divh+F~u.~+89 G Ila II~1 Q II


(4.15)

5. Constitutive Assumptions
As with any theory of material behavior, constitutive assumptions must be
made. These assumptions define the particular type of mixture to be considered.
Among other quantities, constitutive equations must be stated for the mass
supplies, the momentum supplies and the energy supplies. The mass supplies
give rise to chemical reactions and the momentum supplies to diffusion. The
phenomena associated with energy supplies has no classical designation but ari-
ses in a mixture theory which has a separate temperature for each constituent.
It is clear that the constitutive equations must provide relationships between the
various supplies and the 3 91 fields ~ , P~, and aO~, a = 1,..., 9l. It is natural to
consider constrained theories that restrict these fields in some way. For example,
we could require that Z~ and P~ be related in such a fashion that p ldet F 1=0.
I1 tt

This type of assumption corresponds to the case when the ath constituent is
inert. If we place this type of restriction on every constituent then, as (3.1) and
(2.6) imply, the theory is one without chemical reactions. We are interested here
in a mixture theory with diffusion and chemical reactions in which each consti-
tuent is constrained to have the same temperature as every other constituent.
In this special case, it is not essential to use the energy equation for the consti-
tuents (3.7), and, thus, it is not necessary to prescribe constitutive equations for
the partial heat fluxes and the energy supplies. In the list of balance equations we
can replace (3.7) and (3.13) by (3.28) or, equivalently, (4.15). Also, the entropy
inequality appropriate to this special case is (4.10). The 29l + 1 fields one would
naturally consider here are the 91 deformation functions, the 91 density fields,
and the temperature field. In an obvious notation, these fields are
x = z a~ ( Xa , t),
p =P~(X, t) (5.1)
and
0 = 0 (x, t),
where 0 is regarded as a function of spatial position and time. It turns out to
be computatiortally more convenient to replace the 9l density fields P~, a =
1..... 91, by 91 fields ~ defined by a

F~(X, t) = P~ (X, t) det IGRAD Z~(X, t)[. (5.2)


Q r

If we denote the value of F~ by ~, then (5.2) can be written in the simple notation

= p IdetF I. (5.3)
Thermochemistry of a Reacting, Diffusing Mixture 109

By use of (5.3), equation (3.1) can be written

a a
(5.4)
where

fl=ldetFl~.
a tl i1
(5.5)

Clearly, if there are no chemical reactions the function F~ is independent of time


and 7 can be identified as the density of the a tu constituent in the reference con-
a

figuration r. As was mentioned earlier, Z~, a = 1..... 9l, is assumed to be of


a a

class C a. We also shall assume that O is of class C 2 and F~, a = l .... ,91, is of
class C 2"
A thermochemical process for the mixture with a single temperature is a set
of 8 91+4 functions whose values are

a a a a a '

r/=q(x,t), q=q(x,t), r=r(x,t) and O=O(x,t),


a = l , . . . , 9l which satisfy (5.4), (3.3) and (4.15) for a = l .... ,91. As constitutive
assumptions, it is felt that for the case under discussion one should postulate
that ~, T,/~, a a ~/'
a a/~' a = 1,... ' ~ , and q and r/are determined by ~ ~ F~ ' b = 1..... 91,
and 6}. Formally, we can write 1

( ~ , T, l~, ~ , fl~,q, q) = f ( O , Z,, F~). (5.6)

The special case of (5.6) that we shall investigate here is the following

(5.7)
(.*, b , f,F
where f is defined on some open subset of (0, oo) x q / x J X e ~ x S ~ x q / ~ x
(0, oo)~ x ~ ,
g = grad/9 (x, t),

G=GRAD F (X, t) (5.8)


b b b
and
d = G R A D r~(X, t).
b b b

1 For simplicity in writing, all of the various entries in (5.6) have not been shown explicitly.
For example, the appearance of ~gais to be interpreted as representing (1~, ~, 2 .... ~), and the
appearance of ~ is to be interpreted as representing (z~,l Z2...... ~)"
110 R.M. BowEN:

The constitutive equation (5.7) defines a mixture of elastic materials with heat
conduction, diffusion and chemical reactions. If we were to omit chemical reac-
tions, the resulting theory would correspond to the one examined by BOWEN &
WIESE [11]. We shall assume that the f u n c t i o n f in (5.7) and, thus, its component
functions are of class C z. Weaker smoothness assumptions can be stated, but
this one is sufficient for our purposes. In order to simplify the notation, the
component functions of the function f in (5.7) shall be given the same symbol
as their values. For example, (5.7) implies that

~=~(O,g,F,G,x,~,d).b b (5.9)

It also follows from (5.7) along with (4.11) and (4.12) that

r=K(0, g,V,b t,b d). (5.10)


For the sake of simplicity, the constitutive functions are taken to be independent
of X, b = 1. . . . . 9/. When there are no chemical reactions j~ =0, a = 1..... 9l, and
a

(5.4) integrates, for all X, to


r

a~=pa~, (5.11)

where p~ is the density of the a th constituent in its reference configuration. Since


II

these densities are regarded here as known constants, we see that when there are
no chemical reactions the dependence on y and d, b = 1..... 91 can be eliminated
from (5.7). b b
An admissible thermochemical process for the mixture of elastic materials
with a single temperature, heat conduction, diffusion and chemical reactions is
a thermochemical process that is consistent with the constitutive assumptions
(5.7). For every choice of the functions ~ , F~, b = 1.... ,91, and O such that
(5.4) is satisfied, there exists a unique admissible thermochemical process. In
order to confirm this assertion, we see that given the functions ~ , F~, b = 1,..., 91,
and 61 such that (5.4) is satisfied, we can use the constitutive equations (5.8) to
compute (~, aT, ~, ~), a = 1..... 91, q and r/. With this collection of information,
we can compute b,
a
a = l .... ,91, from (3.3) and r from (4.15). The calculation
of r also requires the use of (4.1 1), (4.12) and (5.5). The information just calculated
yields an admissible thermochemical process. Its uniqueness should be evident.
The discussion thus far does not make use of all the equations of balance or
of the entropy inequality. We have the choice at this point to regard these remain-
ing equations as restrictions on the constitutive functions or as restrictions on
the independent variables in the domain of the constitutive functions. Following
and extending the original idea of COLEMAN & NOLL [22], I shall choose the for-
mer and require that (3.2), (3.10), (3.1 1), (3.12) and the entropy inequality (4.3)
be satisfied for every admissible thermochemical process. With the exception of
(4.3), all of these equations are easily satisfied by including their restrictions as
Thermochemistry of a Reacting, Diffusing Mixture 111

part of the defining constitutive assumptions. For example, we can satisfy (3.2)
by prescribing a constitutive equation for each j, in (3.30) and then, if needed,
we can compute the mass supplies from (3.30) and j~, a = 1..... ~ , from (5.5).
a

The interesting restrictions on the constitutive functions will arise from the
inequality (4.3). We shall therefore assume in the following section that (3.2),
(3.10), (3.11) and (3.12) are satisfied.
Since admissible thermochemical processes are generated by the functions
aZ~, F~, a = 1.... ,9~ and O which satisfy (5.4), (4.3) must be satisfied for every
choice of these functions. For ease of reference below, the set of 29~+ 1-tuples
(Z~, F~, O) such that each element generates an admissible thermochemical pro-
cess is denoted by d . Naturally, this set is assumed to be nonempty. In addition
the following assumption is made about the elements of me.
Assumption. Given an arbitrary point x ~ an arbitrary time t o and an arbi-
trary element (a/~, F~, O) of d such that at X ~ =Z~-1 (x ~ t), a = l , . . . , ~ , the
~0
argument o f f is (0 ~ gO,a aF~ aG~ ax , aY~ dO).a It is assumed that (az*, F*, O*) is
also in ~r and satisfies the following conditions:
x*(X,t)=x ~+ x~176176 2+ ( F ~ + A ( t - t~ ( X - X ~
+ (fo + to)) t(x_ xo) | ( x - xo)l (5.12)
t ( x - x ~ | ( x - x ~ | ( x - xo)1,
aF* (X, t~ ~ + d~ 9(aX- aX~ + 89(S(X- aX~ 9(aX- X ~) (5.13)
and
O*(x, t ) = 0 ~ +c(t-to)+(g~ (x-x~ (R(x-x~ (x-x~ (5.14)
for all (X, t) in A/'(X ~ x [t ~ t o + 6], (x, t) in JV(x ~ x [t ~ t o + 6], and all (a, A, B,
E, c, k, R, S ) i n q/~ x Sa~ xSe~ xSe,m x ( - oo, o o ) x q / x 6 a x6a',where.N'(X ~ is a
neighborhood of X
a
~ JV'(x ~ is the corresponding neighborhood of x ~ and 6 is a
positive number. The quantities .N'(X~ Jff(x ~ and 6 are selected such that
(aX*, F*, O*) always yields values in the domain o f f .
Equation (5.13) can be regarded as an initial condition which the differential
equation (5.4) must satisfy whenever (5.12) and (5.14) are given. Physically the
value of F, is the density of the a th constituent measured per unit volume of the
reference configuration. Among other things, the above assumption implies that
at t o these densities can be altered by various choices of the symmetric linear
transformations S,a
a = 1..... ~. Originally COLEMAN & NOEL [22] considered a
case where initial conditions on the independent fields did not enter into the
argument and, thus, were not a part of the concept of a process. COLEMAN &
GURTIN [24] were the first to consider a case in which an initial condition entered
112 R . M . BOWEN:

into the definition of a process. Thus when they required the entropy inequality
to hold for every process it was implied that a certain initial condition could be
freely selected. In this work the definition of a thermochemical process requires
that the elements of d satisfy (5.4) and, thus, associated with a given element
of ~1 there is a definite initial condition on ~ , a = 1..... 9l. The above assumption
simply extends COLEMAN & GURTIN'S point of view and assumes ~ ' is large
enough to allow functions which satisfy initial conditions such as (5.13).
The restrictions imposed by the entropy inequality are not the only ones the
constitutive functions must obey. The axiom of material frame-indifference
requires that an admissible thermochemical process remain admissible after a
change of frame defined by
x*=c+Q(x-o), (5.15)
where c is a time dependent element of d~ and Q is a time dependent orthogonal
linear transformation. The axiom of material frame-indifference requires that if
(5.7) is valid, then the following equation is also valid:

(5.16)

Equation (5.16) must be satisfied for all (0, g, F, G, •, 7, d)


b in the domain o f f
b b b
as well as for all vectors ~, all orthogonal linear transformations Q and all linear
transformations Q such that QQr=_(QQr)r. Of course, in (5.16) we have
assumed that the domain o f f is large enough to insure that f is defined on the
argument shown. A typical solution of (5.7) and (5.16) is

(y. ea - . ra aF . a c ac
L e-.
a c
q.,)
T g,f f,aRAD C,F (x-x),
' ' ~, d), (5.17)
t~ c b b bo

where
C=FrF (5.18)
b b b

is the right Cauchy-Green tensor for the b th constituent and c and b are fixed
integers satisfying 1 < c < N and 1 < b_< N. The most important observation we
need to make at this point is to observe that material frame-indifference requires
that the response functions depend on the 91 velocities Jc,
a a = 1.... ,91, through
9 / - 1 velocity differences. We shall use this fact in the next section.

6. Consequences of the Entropy Inequality


By requiring that the entropy inequality, written in the form (4.10), be satis-
fied for every admissible thermochemical process, we can follow a standard
procedure and deduce restrictions on the constitutive functions. 2 It follows from
2 See, for example, COLEMAN & NOLL [22], COLEMAN & MIZEL [23], COLEMAN & GURTrN [24]
and BOWEN [10].
T h e r m o c h e m i s t r y of a R e a c t i n g , D i f f u s i n g M i x t u r e 113

(5.9) and (2.18)2 that


~p O~u
a 9 a .

=--frO- (0 + g. ua)+ - ~ - . (g + (grad g) u)

+tr~]/-~--{f+GV (x-x)}~+Z-=w-.(x+FF l(x-x))


b\ b bb . b ] b Ox "b bb . ~"
b
(6.1)
+C /07"/_cb~b
| {G+(GRADGIF-'(I-JO}I+Z."
, ,, a~
(~+d.(F-'(+-~)))
07~
+~ . <'
d+(GRADd)F
b b b
_,,
( ax - x v.>>
.
b

If we use (5.4), then (6.1) can be written, after considerable algebra,


07j O~ O~ O~ 07~
, a 9 a . ~ a .x+}-,==_~+y,w_i_..GRADg,,
a ^ a
b b b
r c o~r 0~,~ "a
+tr ELF ~F--s-~ + (x - x) | F[
f, /t..o {f ubr a ~ e 'bx J b /..a

+c~ i N | f) +~-0-.~ "g+-~-" ((gradg).") (6.2)


07-' 07j
+y,(,,-x)-~ f-" ~tfl +f-'
a v ~b
taRAD~
+__~_7 F
a --1
T d+F-,T(GRADdlo_~b
b

If we now substitute (6.2) into (4.10), the result can be written

-
{P'7+-~O-) o~ of

od , , o~, ,^ o~ X
-tr~r-'~pK+v=-~
~ . o , ~ , +2(x-x)
,, , | )
0~ (6.3)
-Z +~-[detF [u-p-L--- "((gradg) a~)
~ )o od
r i OI/s \ 0~
~,..(_' ' - x ) . ~F-irt--~a
OG [GRAD_ff_])+F-ir(GRADd)
r~ b b Od
>~f-O~-
' b b +
114 R . M . BOWEN:

where
(6.4)
and
a~ ~ m
d~
T
f ~ = / ~ _ F - i r O~t [ G I + ~ F - 1 - ~a- - [ G I - - -O~t'I
~-F -1
rd+~^aF-~rd. (6.5)
a a a OF a bb a b (~7 b b
a b a b

As follows from (6.4) and (6.5), the vectorsj~, b = 1..... 91, satisfy the relation
b

2ff=2~. (6.6)

If we use the fact that a~ is expressible as a function of (0, g, F, G, ~, b?, d), then

we can compute the gradient of ~a and form the product - - ~ - . GRAD a~ which
a

appears in (6.3). Finally, if this product is substituted into (6.3), we find that

O~

el

9# ,9# , o# e# 7
- ~-f - - . O~e, F(gradg) b_~_.+EF_lr{b_~_ta]l+2~__~_F_l a+-~ffgj

a~ a~
, .
I (~b[GRADG]~+F-1T(GRADd) b__~__
1
-~F--.'!F- T _~_b[GRADG ] +F-IT(GRADd) h_L > 0

Equation (6.7) must be satisfied for every admissible thermochemical pro-


cess. By an argument 3 which makes use of the last assumption in Section 5, we
can now easily show that in order for (6.7) to be satisfied for every admissible
3 The formal argument is similar to the one used by BOWEN [10, Section 4].
Thermochemistry of a Reacting, Diffusing Mixture 115

thermochemical process the following equations must hold for a = 1.... , ~R

pr/= c30 ' (6.8)

0kut = 0 , (6.9)
0g

0~i = 0 , (6.10)
0k
Q

a~,, =0, (6.11)


~G
o

pK=-F---~F-- | b , , | ~b (6.12)
. a . OF b b Od
a b a a

for all R in 5~

for all Ea in 5:4,


^

--~-b
d" F - , r aS =0 (6.15)
b IJ a

for all S in 5:, and


o

- Z F- F - - . [C,]+---~d+6-wa-~ g~>O.
o,,., \Of o /

In writing (6.13), (6.4)1 and (6.9) were used. Also (6.11) and (6.4)1 were used
in obtaining (6.14). An important observation that can be made regarding (6.16)
is that, in contrast to a theory without chemical reactions, nonzero values of
the gradients G and d, a = 1..... 9l, can cause an entropy production. An examina-
tion of (6.9) through (6.11) shows that

~ ' t i = ~ ' t l ( O , F , ~, b~). (6.17)


a b
116 R . M . BOWEN:

If we use (4.11) and (6.12), we can obtain an expression for the partial stres-
ses. The result can be written

r T = ~ l + e ~ f f + 2 ( ' ,. ~ ~ra~, o~
a a a Or b x-x)| +~b ~-d-| O+" (6.18)
a a b a

Other forms of (6.18) follow by use of the identities

b O (~)_(~_p/p~1)l ' (6.19)


a a

Fa y = P~ ~ - Pa/P~'i ", (6.20)


a a

-
(6.21)

and

The various expressions for Ta are

TT F a~/~ 0 l:~(~u+FO~l~ (6.23)

and
r O~/r 07/ 0~- I
[T=IpFL ~ | , ~r~ + F - ~ L* b -0~-~J " (6.24)

The expression for the moments h~/,


a
a = 1, ..., 9~, can be read off from (6.12) and
(4.14). The expressions for the stress T1 follow from (3.26), (6.23), (6.24) and
(6.18). The results are
.... r OOr o~e[ . . . .
T, = p Z F - ~ F I = Z pF O--~-= Z F ---~--. a-r,. (6.25)
a a, b a a

In obtaining the expression (6.25)1, we have used (6.23) and the fact that

(6.26)
o ~\b~ b %ab/
Thermochemistry of a Reacting, Diffusing Mixture 117

Equation (6.26) is a consequence of material frame-indifference which, as was


/ k

mentioned in Section 5, requires that the vector ~{~u+F


\~ ~ b f f ~ , / ~ depend on

the 91 vectors x, x2 ..... k~ through 91/-1 velocity differences (see (5.17)). Equations
(6.25)2 and (6.25) z follow by a similar argument. This argument also involves
the use of (6.4) and (6.10). In writing (6.25) use has been made of the fact that
TI is symmetric. This condition is automatically satisfied by (6.25) when ~ is
required to satisfy the axiom of material frame-indifference. The proof of this
assertion is an immediate generalization of NOLL'S proof for hyperelastic mate-
rial [18, Section 84]. It is interesting to observe that while the partial stresses
depend explicitly on the mass supplies ~, a = 1,..., 9l, the stress Tt does not.
a

Equations (6.13) through (6.15) place restrictions on ~ and ~, b = 1.... ,91.


b b
However, these restrictions are more complicated than (6.8) through (6.12).
Equations (6.13) through (6.15) do imply certain results which are easy to obtain.
For example, if we set all of the velocities equal to zero, it follows that

~F--~8~"(O'F'~'d.)o(RVb
fl(O'g'Fb'Gb'O'd~))Sg
b =0 (6.27,
o

for all R in 6a,


O~I(O"F'~)'db) ~ -1TffO~(O'g'~b'~'O'~b'~)) }

for all Ea in 5~4 and

for all Sa in ~ . If we differentiate (6.13) with respect to •a and then evaluate the

result at ) = 0 , b = l , . . . , 91, we find


b

R--
a ba
Og
02 e __ (6.30)

a b
for all R in 5P. If we set R=I in (6.30) we immediately obtain a condition on the
a
derivative - ~ - at (0, g, F, G, 0, by, d). Similar results follow by differentiating
(6.14) and (6.15). 4
4 For the case when there are no chemical reactions see BOWEN& WIESE[11, Section 5].
9 Arch. R a t i o n a l Mech. Anal., Vol. 34
118 R.M. BOWEN:

It is worthy of comment at this point that if we had omitted a dependence


on G and d, b = 1, ..., 9~ from this list of independent variables in our constitu-
b
tire functions, we should have obtained additional, but undesirable, infor-
mation from (6.16). If, for this case, we use (6.5), (6.4)and (2.14), equation (6.16)
implies that

for all bGin 5ra and (6.31)

0
b

for all d in ~ . It is easy to show that (6.31) implies that the constitutive equa-
tion ~=~(0, g, bF,x, ~) satisfy, for b # a , the conditions
O~(O,g,F,O,?)
a b b
=0
0F
b

and (6.32)
a~e (O, g,F, O, r)
a b b
=0.
d?
b

Equations (6.32) show that ~ is independent of F and ?, for b4:a, when Jc =0,
a b b b
b = 1 , . . . , ~ . As mentioned by BOWEN [1] and BOWEN • WIESE [11], it is incon-
sistent with classical thermochemistry for (6.32) to be valid except as a restric-
tive special case. This point was also stressed by MOLLER [8].

7. Equilibrium Consequences of the Entropy Inequality


Equations (6.8) through (6.16) are necessary and sufficient conditions that
the entropy inequality (4.10) be satisfied for every admissible thermochemical
process. It is possible to extract additional restrictions on the constitutive func-
tions from the entropy inequality (6.16). These additional results fall into a dif-
ferent class from the previous ones because they are only true for certain values
of the independent variables. It is usually convenient to assign the name equi-
librium to the state in which these special values occur. However, we can formally
obtain these results without injecting at this point the physical issue as to whether
or not the word equilibrium should be used. We shall assume, in addition to
the previous assumptions, that for some element in the domain of the constitu-
tive function of the form (0 +, 0, bF + , 0, 0, ~+, 0) the t e r m -~ --~-fl
a~i ^ is zero when
a

evaluated on this element. We shall discuss later in this section some of the
various possibilities which achieve this result. We shall adopt the notation f +
Thermochemistry of a Reacting, Diffusing Mixture 119

to indicate a functionf evaluated at (0 +, 0, bF + ,0, 0, yb+, 0). If

, I^~
~(o, g,f , f, ,~, ~, #)= -glO . h - 2 ~, "~ , ~ + ~ g)

,-, { 0~e~ 1 .detF_,l u2"~j (7.1)


-~t~+~ -, o o~o

..~0 b Off --~[~]+--~-~ d+4-~-~ r r gl,


then (6.16) shows that
I(0, g, if, G, ~, y, d ) ~ 0 (7.2)
D b o

and, by the above assumption,


l(O+, O, F +, 0, O, ~,+, O)=0. (7.3)
Therefore,
d l(O++~.c~,2a,F++2A,2K, 2m, yb++2~,2k )
d2 b b b =0 (7.4)
~,=0
and
d 2 l(O++2~,2a, F++2A,2K, 2ra,~,++2lA2k)l ~,.,
d2 2 b b ~ ~ h b o la=o~V (7.5)

for all (~, a, A, K, m, ~, kb)in (--co, ~ ) x q / x o ~ a ~ x6e~ x o//~qx ( - - ~ , ~)sa xq/'.


Equations (7.4) and (7.1) imply that
-./o+ ~ + - zq- +~ - c + W > z ~ ~

/O/~+ O/~+ 0/~+ ", (7.6)


b b b +T

where c~t + is the concentration of the a th constituent evaluated at the + state.


Actually, the term ~ f + in (7.6) is zero because (6.6) and (3.10) imply that
b

~ f ^+ = ~ p ^+ = 0 9 (7.7)
Since h

t~2(a___ ~, ^ +~ . or+ / ~

+ _ _ ~ _ L _b_~Z . ,~ + o o ~,, ^ + a or, +


~+ .~,
a b

9*
120 R.M. BOWE~:

(7.6) readily implies that


0 (7.9)

h*=q =-o*Z o~, Og o+Ze*~ oo' (7.1o)


a Q

a'Pl^ *

O~ Ct-~a | r,a r~ ~O~-.( O_~_._[K])=0 (7.12)

for all K
a
in ~ ,

f+=~+= -Z or/a/~*~ (7.13)


a a O~' 02 '
b a

a, =o (7.14)

and
+ 0~+-1
+ F + -*F* =0. (7.15)

In writing (7.8) use was made of the fact that 7ti is independent of g, G and 2.b
Observe that (7.13) is consistent with (7.7) because, as earlier with (6.26), mate-
b
rial frame-indifference implies that --~-~-=0" It is interesting to observe from
fl

(7.10) and (7.13) that, as a consequence of the chemical reactions, the heat flux
and the momentum supplies do not necessarily vanish in the state (0 +, 0, Fb + ' 0,
0, y+, 0). For simplicity, we shall not calculate the restrictions implied by (7.5)
because of the complications it would entail.
There are two natural choices to make in order to achieve the result
O~l* ^ .
~'--~7 ~ =0. (7.16)
tl

The first case one might take is to require, for ~ = 1.... ,9~,

j~(O+, O, F +, O, O, y+, 0 ) = 0 , (7.17)


b
Thermochemistry of a Reacting, Diffusing Mixture 121

then, by ( 3 . 3 0 ) , ( 5 . 5 ) a n d ( 2 . 6 ) , f o r a = 1 . . . . , ~ ,

+, o,U, o, o, o)=o. (7.18)


a u b

The state such that (7.17) is satisfied was called an equilibrium state by BOWEN
[10]. I shall adopt here the nomenclature suggested later by TRUESDELL[16, Chap-
ter 6] and refer to a state (0 +, 0, F~ + , 0, 0, ~+, 0) such that (7.17) is satisfied as a
state of weak equilibrium. Equations (7.9) through (7.15) can be easily specialized
to this case. For example, (7.14) simplifies to

b =0. (7.19)
07 Oy
b a

The meaning of (7.19) becomes clear if we use (3.30) and (5.5) and write ~ ~-~--~b
in the form b

#= - E tr~j~, (7.20)
b ry b a=l
b

where ~ is the chemical affinity for the ~th chemical reaction and is defined by 5

a~ = - ~ m* Idet F I P,~ 8 y x = try(O,F, y, d). (7.21)


a O~ b b
a

By using (7.20), we can write (7.19) in the form


91 .+
X~ aJ~ a~+_0
/ , - - l -- 9 (7.22)
9=1 07'
a

Equation (7.22) shows that in a state of weak equilibrium the 9t dimensional vec-
tor (a~+, ~2+,..., ~ + ) must lie in the kernel of the 91x91 dimensional matrix
- ~ . . - , . As a special case of (7.22), if OJ+~ r is regular then
-,vj

ax(O ,F
b +'~? +, 0 ) = 0 ct=l . . . . . 91. (7.23)

As mentioned by BOWEN [10], classical thermochemistry texts often define


equilibrium by (7.23). Following the nomenclature suggested by TRUESDELL, a
state in which both the chemical reaction rates and the chemical affinities vanish
shall be called a state of strong equilibrium. We have just shown that a weak
equilibrium state is necessarily a strong one if LT -j is regular at (0 +, O, rb '
0,0, ~+, 0).
b

5 When there is no diffusion, it is easy to show that (7.21) reduces to the definition given by
BOWEN[10].
122 R.M. BOWEN:

As a second example when (7.16) is true, we can use the identity (7.20) and
achieve (7.16) by requiring, for ~ = 1..... ~R,
~t +
az(O ,F + ? + 0 ) = 0 (7.24)
b ~b '

in the state (0 § 0, F § 0, 0, y+, 0). In this case (7.14) reduces to

OffI .+
,=27,1---~ J , = 0 , (7.25)
r

where (7.20) has been used. Equation (7.25) shows that in a state such that (7.24)
is true the ~ dm
i ensoio~+
naF
l y vector (j. +,,J2,...,
.+
J+) must lie in the kernel of FL oa~+Ij 9 9
Therefore, if L aay .] is regular, then

j,(O +, O,F +, O, O, ? +, 0 ) = 0 (7.26)


b

wheneverFa~l (724) is true. If we compare (7.25) and (7.22) we see that ~ both j
and L-TfI are regular at (0 +, O, F O, O, ~7+, 0), then

:~(o+, o,U, o, o,~+, o)=o


if and only if (7.27)
o~rO
I\
+' bF +~b7 + , 0 ) = 0
for a = l , . . . , ~.
In the case that is regular at (0 +, 0, ~F + ,0, 0, ?+, 0) our smooth-

ness assumptions and (7.17) insure that there exists a neighborhood J// con-
taining (O+'F+b' 7+'"'1 ' ~+), and, for ~ = ~ + 1 , . . . , 9/, a neighborhood JV(~ +)
containing ? + and a C 1 function ~+: ~ ' ~sV'(~ +) such that for all (0, (, 71..... 2)
in d / ~
j,(O, O,(, O, O, ~..... ~+17'
?+(0, F,
7 ~1 ..... ~Y)..... ~?-+(0, F,o 17'''" ,7),
0)=0~ (7.28)

for ~ = 1, ..., 9L A similar assertion can be made about inverting the ~ equations

a~(0, F, ?, 0 ) = 0 (7.29)
b
9
when L ~ J is regular at (0 +, O, F+, O, O,~+, O) and when (7.24) is satisfied.

Before closing this section, it is worth remarking that in the special case when
there are no chemical reactions the results analogous to (7.9) through (7.15) can
be obtained without requiring that G =0, b = 1..... 9L The results of this calcula-
b
tion can be found in the paper by BOWEN & WIwsE [11].
Thermochemistryof a Reacting, Diffusing Mixture 123

8. Material Symmetry
For a mixture with chemical reactions and diffusion, the ideas of material
symmetry are particularly interesting since the relationship between the density
fields and deformation fields is such that it is no longer natural to regard the
isotropy group as a subgroup of the unimodular group 6. As we shall see below,
it is this fact that allows the results presented in Sections 6 and 7 to be specialized
directly to the case of a mixture of chemically reacting fluids. The isotropy group
J for the b th constituent in a chemically reacting mixture of elastic materials
b
with diffusion and heat conduction is defined to be the set of automorphisms
such that if
(~, T , ~ , M , fl, b b (8.1)
then
(~, T, ~, ~ , ~, ... , ~ [det HI, ... , ~ , q , q )

= f ( O , g , F , . . . , F H ..... F,G, C(G|174 ..... G,~r 7 ..... 7[detH[ .... (8.2)


1 ""' b 1 b

~ , , d , . . . , I d e t H I H T d . . . . . d).
91

Equation (8.2) is required to hold for all values of the argument (0, g, F, G, :~, ~, d).
b b

The quantities V and d in (8.2) transform under the mapping F--*FH in such a
b b b b

fashion to insure that p ~ p and ~'--* Z. This observation follows from (5.3) and
b b b b
(5.5). If the b th constituent were inert, then ~ = 0 for all time t and, as follows
from (5.11) and (5.3), b
p [detF [ =PK. (8.3)
b b

In this case, the above transformation rules require that [det HI = 1 which implies
the usual condition that J i s a subgroup of the unimodular group. As suggested
b
earlier, the special case of a mixture of chemically reacting fluids is especially
interesting. The defining symmetry for the b th constituent to be a fluid is taken
to be J = J.Y. Thus any nonsingular linear transformation H is allowed in (8.2).
b
For the circumstances where every constituent is a fluid, we can select b = 1 and
H = F -1 in (8.2) and find that
I

(8.4)
.... .... . . . . .

where
-1)
and (8.5)
s = [detF -1 IF -1 r d = [detF- 11 grad 7.
a a a a a a

6 T R UE S DE L L 8s N O L L [18, Section 31].


124 R . M . BOWEN:

If we now substitute (8.4) into (8.2) and repeat the above argument for b =
2, 3 ..... 91, we find that

(~T' p'I'~I~' q't'l)=f(Og'ttI'Y~' P' ~' )a'a (8.6)

Equation (8.6) is a necessary condition that follows from (8.2). Conversely, (8.6)
is a sufficient condition that insures that (8.2) is satisfied for ~r a, b = 1..... 9l.
It follows from (8.6)that there exists a function f such that b

(8.7)

or, equivalently from (8.5) and (5.3), there exists a function f such that

(e, r,
a a
~, ~,
a
~, q, ~)=i(o, g, ~, ~,D bp,.O,
a D
(8.8)
where
e = grad p = grad(y Idet F-11 = s - p C12 M . (8.9)
b b ) v b D

If we now incorporate the restrictions on f and f implied by material frame-


indifference, it is easily shown from (5.16) that the axiom of material frame-
indifference is satisfied if and only if

(y. T.~ q. ,)=f*(O.g.M.b bu'


(8.10)
=f**(0, g, M,bbbu'p, el,
where
u=x-x, (8.11)
bb b b

b is any fixed integer such that 1 < b < 9 1 and f * and f * * are isotropicfunctions.
We can now rewrite certain of the results of Section 6 for the special case of a
mixture of fluids. It follows from (8.10)1, (6.17) and (6.8) that

7tz= ~J* (0, p, ~) (8.12)


and
O~r*(0, p, g). (8.13)
P~/= 00

Equations similar to (8.12) and (8.13) follow from (8.10)2; however, the occur-
rence of the independent variable e = g r a d p complicates matters because, from
(8.9)3, ~** would now depend upon M, 5 = l , . . . , 91, in a special way. For
simplicity, we shall base all of the following calculations on (8.10)1. By a change
Thermochemistryof a Reacting, DiffusingMixture 125

of variables, it can be shown that


T
b
Fa _ _
8Fa

(8.14)

8~*(o,~ g, M,~ ~u' f, ~) 8~*(0, g, M,~ ~' ~, ~)


a a

F 8d = l d e t ,F - q - - 8s ' (8.15)
a a

8y(O,g,F, G.~.. y,d) 8 y * ( O . g . M . u . p,~)


(8.16)
a e~b eb

and
8~c (o, g, ~F'o,~ ~,~~' ~) ,-,~ ~ 8e*c(o, g, ~,& p,;)
]detF- 11 _ _
(8.17)
c
a e~b eb

Therefore, from (6.18), (6.4), (8.12) and (8.14) through (8.17)

rT = ....
~"L - p - -8~r,,.
~ - l - ~ - -,~e,
| s-s. 8~x I
i1

a~ts (8.18)
8~
b
+2b Zr (a-.~),,
a a bb
| ca b
Er (~o-L)--ff~| 8u"
c~b cb r b cb

The inner part of the stress Tx is given by

(8.19)

In terms of the free energy densities, (8.18) and (8.19) can be written

Off .8~

(8.20)
oq,
b ~
o4,
9
o~b
b ca oa rJb e,b
c#b Cb r b Cb
126 R . M . BOWEN:

and
aq,
(8.21)
a o, b b \

The chemical affinity, defined by (7.21), becomes

a ,~'= - ~ m QP~~ d~tdp= -P~ m '~e~~' d@t


ap , (8.22)
a a
(l r

where (2.20) and (3.31) have been used in obtaining (8.22)2. In the theories of
reacting fluids discussed in the textbooks 7 on irreversible thermodynamics the
chemical affinity is usually expressed in terms of the chemical potentials. The
generalization of this classical expression is easily obtained if (8.18) is solved
for - ~ p and the result, along with (4.11), is substituted into (8.22)1. However,
r
this calculation is somewhat artificial since there is no natural relationship
between the chemical potentials and the chemical affinities.
The above formulas for TI and try reduce entirely to the classical expressions
under some natural approximations. For example, if we were to investigate a
theory in which the departure from equilibrium (either weak or strong) is small
then the fact that f * in (8.10)1 is isotropic allows us to approximate Tr and
V~I by functions which, to the first order, are independent of ~, b = 1,..., 9l.

Acknowledgment. The research reported here was supported by the U.S. National Science
Foundation under Grants GU-1153 and GP-9492.

References
1. BOWEN,R. M., Arch. Rational Mech. Anal. 24, 370--403 (1967).
2. GREEN, A. E., & P. M. NAGHDI, Intl. J. of Engr. Sci. 3, 231--242 (1965).
3. CROCHET,M. J., & P. M. NAGHDI, Intl. J. Engr. Sci. 4, 383--402 (1966).
4. GREEN,A. E., & T. R. STEEL, Intl. J. Engr. Sci. 4, 483-- 500 (1966).
5. STEEL,T. R., Quart J. Mech. and Applied Mech. 20, 47--72 (1967).
6. STEEL,T. R., Intl. J. Engr. Sci. 5, 775-- 790 (1967).
7. MILLS, N., Intl. J. Engr. Sci. 4, 97-- 112 (1966).
8. MOLLER,I., Arch. Rational Mech. Anal. 28, 1 -- 39 (1968).
9. INGRAM, J. D . , & A . C . ERINGEN, Intl. J. Engr. Sci. 4, 289--322 (1967).
10. BowE~, R. M., J. Chem. Phys. 49, 1625-- 1637 (1968).
11. BOWEN,R. M., & J. C. Wt~E, Intl. J. Engr. Sci. (forthcoming).
12. TRUESDELL,C., Rend. Accad. Lincei (8) 22, 33--88, 158--166 (1957).
13. TRUESDELL,C., & R. TOUPIN, The Classical Field Theories. Handbuch der Physik, Band III/1
(edited by S. FLOOOE). Berlin-G6ttingen-Heidelberg: Springer 1960.
14. KELLY, P., Intl. J. Engr. Sci. 2, 129--153 (1964).
15. ERINGEN,A. C., &: J. D. INGRAM, Intl. J. Engr. Sci. 3, 197--212 (1965).
16. TRUESDELL,C., Rational Thermodynamics, A Course of Lectures on Selected Topics. New
York: McGraw-Hill (in press).

7 See, for example, DE G g o o r & MAztm [25, Chapter 2].


Thermochemistry of a Reacting, Diffusing Mixture 127

17. GREEN,A. E., & P. M. i'q'AGHDI,Intl. J. Engr. Sci. 6, 631--635 (1968).


18. TRUESDELL,C., & W. NOEL, The Nonlinear Field Theories of Mechanics. Handbuch der
Physik, Band III/3 (edited by S. FLi3GOE). Berlin-Heidelberg-New York: Springer 1965.
19. NOEL,W., The Axiomatic Method, with Special Reference to Geometry and Physics, 266--
281. Amsterdam: North Holland Co. 1959.
20. GREEN,A. E., & P. M. NAGHDI,Arch. Rational Mech. Anal. 24, 243--263 (1967).
21. BOWEN,R. M., Arch. Rational Mech. Anal. 29, 114--124 (1968).
22. COLEMAN,B. D., & W. NOEL, Arch. Rational Mech. Anal. 13, 167-- 178 (1963).
23. COLEMAN,B. D., & V. J. MIZEL,J. Chem. Phys. 40, 1116--1125 (1964).
24. COLEMAN,B. D., & M. E. GURTtN,J. Chem. Phys. 47, 597--613 (1967).
25. DE GROOT,S.R., & P. MAZUR, Non-Equilibrium Thermodynamics. New York: Inter-
science Pub1. 1962.
Rice University
Houston, Texas
(Received February 18, 1969)

You might also like