M.A. A Biot - Fluence
M.A. A Biot - Fluence
M.A. A Biot - Fluence
New Variational-Lagrangian
Irreversible Thermodynamics with Application to
Viscous Flow, Reaction-Diffusion,
and Solid Mechanics
M. A. BIOT
I. Introduction ........................................ 2
II. Restructured Thermodynamics of Open Systems and the Concept of
Thermobaric Transfer. ................................. 6
III. New Chemical Thermodynamics .......................... 9
IV. Homogeneous Mixtures and Reformulation of the Gibbs-Duhem
Theorem .......................................... 13
V. Nontensorial Virtual Work Approach to Finite Strain and Stress ..... 16
VI. Thermodynamic Functions of Open Deformable Solids. ........... 19
VII. The Fluence Concept. ................................. 21
VIII. The Nature of Entropy Production and Its Evaluation ............ 23
IX. The Principle of Virtual Dissipation ........................ 30
X. General Lagrangian Equations ............................ 3.5
XI. Dynamics of Viscous Fluid Mixtures with Reaction-Diffusion and
Radiation Pressure. ................................... 36
XII. Dynamics of Solids with Elastoviscous Stresses and Heat Conduction,
and Thermoelasticity .................................. 44
XIII. Inhomogeneous Viscous Fluid with Convected Coordinates and Heat
Conduction. ........................................ 51
XIV. Lagrangian Equations of Heat Transfer and Their Mechanical
Interpretation, and a Mass Transfer Analogy .................. 56
xv. Deformable Solids with Thermomolecular Diffusion and Chemical
Reactions. ......................................... 59
XVI. Thermodynamics of Nonlinear Viscoelasticity and Plasticity with
Internal Coordinates and Heredity ......................... 62
XVII. Dynamics of a Fluid-Saturated Deformable Porous Solid with Heat and
Mass Transfer. ...................................... 64
XVIII. Linear Thermodynamics near Equilibrium .................... 68
XIX. Linear Thermodynamics of a Solid under Initial Stress. ........... 13
I
Copyright 6 1984 by Academic Press, Inc.
All rights of reproduction in any form reserved.
ISBN O-12-002024-6
2 M. A. Biot
I. Introduction
d%=-pdv+xEkdMk+dh. (2.1)
k
Ek = (2.2)
New Variational-Lagrangian Irreversible Thermodynamics 7
where pi, pi, T’, and d$ are, respectively, the pressure, density, temper-
ature, and specific entropy differential of the mass dM” along the path of
integration. The value of Ek includes the work of extraction of dMk from
Csk and injection into Cr , and T’ dfi is the heat injected into dMk at every
step of the thermobaric transfer.
Similarly, the increase of entropy of the subsystem Cr + c Csk is
PLT
3 dY = T Sk dM” + dhlT, Sk =
I
P,~~o
dG . (2.3)
We called pk the injection pressure and Ek and Sk, IXSpeCtiVdy, the in&X-
tion enthalpy and injection entropy. From Eq. (2.2) we derive the impor-
tant differential relation for each substance
-&(hkT - p) = $, (2.6)
The subscript urn indicates that the derivative is for u and Mk constant.
Similarly, we obtain the relations
(2.9) *
(2.10) c
(2.12)
The derivatives in Eqs. (2.8) and (2.12) are obtained from relations
P = P(U, Mk, 0, (2.13)
dQ=--pdu+~+kdMk+TdY, (2.17)
k
New Variational-Lagrangian Irreversible Thermodynamics 9
where
& = Ek - TFk (2.18)
is the convective potential (Biot, 1976a, 1977a). Relation (2.17) is analo-
gous to the Gibbs equation with & replacing the chemical potential. In
contrast with the classic procedure, Eq. (2.17) is not used to define & but
constitutes a theorem. In addition, &, Ek, Sk do not involve undetermined
9 constants as in the classic case. Note that these undetermined constants
are not eliminated by taking differentials, because we then obtain
_ d& = d& - T d$k - Sk dT, (2.19)
where, if we follow standard procedures, an undetermined constant still
remains for nonisothermal transformations in the coefficient Sk of dT. This
difficulty was already recognized by Gibbs himself (Gibbs, 1906) as well
as by others (see, e.g., Hatsopoulos and Keenan, 1965).
Gibbs’s paradox is also eliminated, as shown by Eq. (2.3), because for
identical substances the injection pressures I)k and hence also the injection
entropies Sk are the same, with the result that in Eq. (2.3) entropies
become additive. It is important to note that, when two components
become identical, Eq. (2.15) for the injection pressure in terms of the
molar fraction loses its validity.
This is consistent with the physical fact that in that case a semiperme-
able membrane loses its ability to distinguish between components.
position of the cell remain unchanged, its state and hence its pressure and
volume do not vary. The constancy of fhe temperature is obtained by
injecting into the cell a quantity of heat hr de. This defines the intrinsic
heat of reaction hT. How this is related to standard definitions can be
shown by reinjecting into the cell the products of the reaction either at
constant volume or at constant pressure, with T constant. We write
hpr = ET + ck vk h;r,
where huT and hPT are the standard heats of reaction at constant tempera-
ture and, respectively, at constant volume and pressure. The heat of
mixing hiT at constant T and u is given by Eq. (2.12) and hiT denotes the
heat of mixing at constant temperature and pressure. By a method similar
to the previous derivation of h$T, we showed (Biot, 1982a) that the value
Of h;T iS
(3.3)
The derivative ~/cJTis for constant pressure and composition. For a per-
fect gas mixture, substitution of the value for& from (2.15) yields hiT = 0.
Hence in this case hpT k = 0, i.e., the standard and intrinsic heats of
reaction coincide.
Note that the intrinsic heat of reaction is more representative of chemi-
cal properties than standard concepts, because it does not involve the
heat of mixing or external work. For this reason its value is obtained by a
very general formula relating the two heats of reaction for two different
states. Consider a rigid cell in state 1 with values pr’, T, and another in
state 2 with values pp’, T2. A reaction d[ occurs in cell 1 and -d[ in cell 2,
while products of reaction are extracted from 1 and injected into 2 by
thermobaric transfer. The temperatures of the cells are kept constant by
injecting amounts of heat hg’ de and -h’,2’ d[, respectively, into each
cell. Conservation of energy implies the relation (Biot, 1977a) generalizing
Kirchhoff’s formula
(3.4)
Hence if we know the intrinsic heat of reaction for a single state, we may
derive its value (as well as the standard values huT and hpT) for any other
state of arbitrary composition and temperature.
New Variational-Lagrangian Irreversible Thermodynamics 11
+zEkdMk+dh 9
k
with
These results are obtained by adding to Eqs. (2.1), (2.3), and (2.5) the
terms due to the masses Vkd[ transferred and the heat hFq d[ injected into
Cch. We have also taken into account the fact that Cch is in equilibrium, so
no change of entropy occurs in the system due to the chemical reaction in
Cch .
The differentials (3.5) and (3.6) may be integrated from a given initial
state, yielding values
dQ=-pdv-Ad.$+~~kdMk+TdY, (3.12)
k
where
h”Tq
= c
k
t+(/:TerdEk + F,(O)), (3.15)
A = -2 vkpk, (3.17)
k
where
dQ=-_pdv-~A,d~p+~~kdMk+TdY. (3.22)
P k
s;q =
and dmk = vk d.f + dMk. For simplicity we assume a single reaction, but
the results may be readily generalized to multiple reactions. We start with
a cell of zero volume. Its volume u is then increased gradually while
maintaining constant the pressure p, the temperature T, and the concen-
trations Of the VariOUS substances. Hence, PkEk and Sk ah0 remain con-
stant. The heat that must be injected is
dh = 2 hiTdmk, (4.5)
k
with
During the transformation considered here, the value of hiT remains con-
stant, hence also the values of %k and yk as well as p and qeq. Thus
integration of Eqs. (4.6) and (4.7) yields
(4.9)
These values are the energy and entropy of a homogeneous cell contain-
ing masses mk = vkt + Mk of each substance.
New Variational-Lagrangian Irreversible Thermodynamics 15
For a pure gas mixture hET = 0, and without chemical reactions (< = 0)
we obtain
(4.11)
k
showing that in this case the enthalpy of the mixture is the sum of the
enthalpies of the individual components, and likewise for the entropy
(Biot, 1982a).
We note again that for identical components the injection pressures pk
are the same, hence also the values of Sk, so that Gibbs’s paradox is
avoided.
To reformulate the Gibbs-Duhem theorem in the present context, we
differentiate Eqs. (4.9) and (4.10). Taking into account the values of these
differentials from Eqs. (4.6) and (4.7), we derive
T mk dYk = 0. (4.13)
u dp = ; mk(d%k - T dYk)
where
9’ = ck mkYk. (4.17)
entropy Yc instead of the chemical potentials & and the entropy used in
the classical form. The present formulation avoids the basic difJiculties of
the classical treatment, which involves undetermined constants in pk,
and in the entropy.
Expressions (4.9) and (4.10) are readily generalized for multiple reac-
tions as
Y=~Y&tp+~Ykmk, (4.19)
P k
where
Xf = R+j. (5.2)
It is then followed by another transformation,
Xi = (6ij + &ij)Xj’ 9
(5.3)
where the coefficients are not independent but satisfy three constraints,
so that they contain only six degrees of freedom. For example, we may
adopt the three constraints
Eij = &ji (5.4)
and call eij the six finite strain components defining the deformation (5.3).
In this case, the transformation (5.3) is such that principal directions of
strain remain invariant (Biot, 1965b).
Other choices may be made. For example, the transformation (5.3) may
be chosen so that the material on the xi axis remains on that axis while the
material in the xix{ plane remains in that plane. In this case, z21 = &31= ~32
= 0, and the six remaining coefficients define the strain components.
Other definitions may be chosen suitable to the particular physical proper-
ties of the material (Biot, 1973a).
For the method to be complete we must be able to express the six strain
components cti in terms of the nine coefficients aij. We have shown that
this can easily be done to any order by a systematic procedure (Biot,
1965b). For example, to the second order, for the choice (5.4), we have
1 1
Eij = eij + j (f?ipOJpj + &jpWpi) + - WipWjp, (5.5)
2
where
1
eij = i (ati + Uji), Oij = 5 (au - Uji). (5.6)
1
~12 = e12 + 7 a2lh22 - ad. (5.8)
with six coefficients rti defining stress components conjugate to Ed. The
quantity 6W is a physical invariant, but the factors rij and 8~ need not be
tensors. The indices i and j indicate a summation to all six independent
terms whatever their nature. If cij is expressed as a function of ati, we may
write
(5.11)
where
TV = rFp &Ida, (5.12)
are the nine components of the Piola tensor. We may also consider the
stress aij per unit area after deformation, referred to the initial axes xi. A
virtual transformation after deformation may be written in terms of final
coordinates Xi as
axi = Fj aa.. (5.13)
U’
where 6Zij are nine suitable coefficients to be determined. The unit cube
after deformation has become a parallelepiped, and the virtual work of the
forces aij in the virtual transformation (5.13) is
6W = Jq Gi,, (5.14)
Since the virtual work (5.14) must vanish for a pure rotation %ig = -Ssl,,
the tensor oV is symmetric and satisfies the three relations
Uij = Uji 3 (5.20)
which also expresses the conditions that the moments due to the nine
stress components Tti vanish.
Substitution in Eq. (5.19) of the values (5.12) of Tij yields
Cri = (IIJ)~p~Cjk d$,/&li~. (5.21)
This expression is valid for all definitions of r@,,, and the summation is
extended to all six components of the stress 7PVas defined above (Biot,
1981).
For a continuous deformation field, we write Xi = xi + Ui(x/), where the
displacement Ui is a function of the initial coordinates Xi. The local infini-
tesimal transformation
dXi = (6ij + auilaxj) dxj (5.22)
is the same as Eq. (5.1) provided we write
dUildXj = au. (5.23)
With ag representing the nine displacement gradients, the foregoing defi-
nitions of stress and strain remain valid as local values for a nonhomoge-
neous deformation.
The collective energy and entropy per unit initial volume of a deforma-
ble solid based on virtual work and nontensorial concepts were derived
earlier (Biot, 1981). The results may be readily extended to an open solid
with chemical reactions by proceeding as in Section III. We consider
initially a cube of unit size. A homogeneous deformation is defined by the
20 M. A. Biot
six strain components aij under the corresponding stresses TV. The solid
contains substances in solution. The increment of mass of each substance
is
dmk = Yk ds$ + dMk, (6.1)
where Vkde is due to a chemical reaction and dMk is the mass acquired
from an external source. For simplicity we assume a single reaction,
because results are readily generalized for multiple reactions. An infini-
tesimal change is defined by dq, dt, dMk and the temperature change dT.
The corresponding changes in energy and entropy are obtained by ther-
mobaric transfer and their values are derived from Eqs. (3.5) and (3.6) by
replacing -p du by rij 8~~. This yields
Fl”T”
dF;+ydt+cFkdMk+$.
eq k
The terms hiT dq represent the heat that must be injected when applying
a deformation dq under the constraint dmk = dT = 0. Similarly, h$ is the
heat of mixing, where hfT dMk is the heat injected when adding a mass
dMk under the constraint dsti = dT = 0. The heat capacity C,, is for de, =
dmk = 0. We may show that h$,T and h$ may be determined without
calorimetric measurements when we know the equations of state
rij = r&&u, mk, T), (6.5)
i.e., when we know the stress rij and the injection pressures pk as func-
tions of the strain F~, the masses mk of each substance added by chemical
reaction and transport, and the temperature T. From the conditions that
Eqs. (6.2) and (6.3) are exact differentials, proceeding exactly as before in
deriving Eqs. (2.8) and (2.12), we obtain
h$r = - T(c%,lc3T),, , (6.7)
where the subscripts indicate that the partial derivatives are for aij and mk
constant.
New Variational-Lagrangian Irreversible Thermodynamics 21
d~=qd~G-~Apd&,+~~~dMk+TdY. (6.11)
P P
(7.2)
by (7.1), where i6lf represents the more general mass flux. Hence Eqs.
(7.1)-(7.3) retain exactly the same form with this generalized definition of
the mass flux &l: and mass fluence M$. In this case, however, Zt%fis not a
Cartesian vector. We denote by &flk the corresponding Cartesian compo-
nents, i.e., tiik represents the mass flux per unit area in the space Xi. An
initial surface A is transformed into A’ by transformation (7.4). The mass
flux through it may be expressed in two different forms as
IA n;r~ni dA =
I A’
n;lj”C~inr dA’, (7.5)
J’ = detIdxilZj/. (7.6)
Obviously, in the second integral of Eq. (7.5) the integrand k;Cji is the
Cartesian component &fIk of the mass flux in the space Xi. Hence
nj,!” = ~Q!c!. (7.7)
J JI’
This relation may be given another useful form by considering the linear
transformation of dfi into dxi and solving for dfi. We find
axilaxj = CjiIJ’. (7.8)
Hence
n;r,!” = Jln;rjk afi/ax.
(7.9)
J’
These results show the mass flux, as defined above for a deformable
coordinate system, to be a relative contravariant vector (Sokolnikoff,
1951, see pp. 58, 72). We shall refer to &ff and Mf, respectively, as the
contravariant mass flux and the contravariant mass jluence.
Similarly, consider the local Cartesian heat flux BI at Xi and a contrava-
riant heat flux Hi defined as the rate of heat flow across a transformed
surface initially equal to unity and initially normal to the Xi axis. They
satisfy the same relations as Eqs. (7.9) and (7.10), namely
New Variational-Lagrangian Irreversible Thermodynamics 23
The time integral of I!Ziis the contravariant heatfluence Hi. We may write
a conservation condition analogous to Eq. (7.3) as
h = -aHiIaXi, (7.13)
where h is the heat acquired by a deformed element of unit initial volume.
The entropy flux across an area is defined as (Biot, 1977b)
(7.14)
Its time integral is the entropy fluence Si. They are either Cartesian or
contravariant components.
The fluence fields Mf are state variables, because they determine Mk by
Eq. (7.3). However, the fluence fields Hi and Si are not state variables.
Other fluence fields that are state variables may be introduced by put-
ting
9 = -aStldxi, % = -aSd:laxi, (7.15)
where ST is called a pseudo-entropy fluence and ST a pseudo-energy
jluence. Using such concepts, the state of a system may be partially
described by fluence vectors as further clarified below in several applica-
tions.
(8.4)
where
Hence 8s*TMis the virtual entropy produced per unit initial volume. It has
three terms. The first is due to the masses convected 6Mf, the second is
due to the heat 6hP produced by the irreversibility, and the third is due to
the heat flow 6Hi across a temperature gradient.
It remains to evaluate the heat produced 6hP. This is obtained by con-
sidering energy conservation. Using the value (8.2) of ah, the variation of
energy (2.1) per unit initial volume with 6v = 0 is
&?L=~EkBMk+6hP-&6Hi. (8.8)
I
Substitution of Eq. (8.3) for 6Mk and integration over a domain fl yields
New Variational-Lagrangian Irreversible Thermodynamics 25
(8.10)
where
(8.13)
It should be understood that these relations also depend on the local state
variables, although this is not formulated explicitly. They embody the
irreversible kinetics of the system and may be obtained either experimen-
tally or theoretically from molecular kinetics. With the values (8.14),
26 M. A. Biot
T ~~~~~~= F G!il~”
6M,k + 3:’ 6Si, (8.16)
where
aFk(h/, Sj) = GRf- TFk%T, (8.17)
~~T(n;rj’, Sj) = Twit. (8: 18)
The present derivation of entropy production is fully general and does not
assume linearity or any dependence on Onsager’s (1930, 1931) principle.
For example, we may include nonlinear diffusion of a non-Newtonian
fluid through a porous medium.
In the case of a linear dependence of ‘%T on the fluxes and in the
absence of a temperature gradient we obtain from (8.14) the linear rela-
tions
%T(n;r,l, Ejj)= 0, (8.19)
where Hi is the coupled heat flux due solely to the mass flux. If we solve
these equations for fii, the coefficients of fi,” constitute what is generally
called the heat of transport.
Consider now the entropy produced in a primary cell by a chemical
reaction. This is the entropy increase of a cell for du = dMk = d% = 0. In
this case Eq. (3.22) is applicable and is written in variational form as
where 8s*Chis the entropy produced by the reactions S&,. The values of
the affinities are functions Of u, mk, T,
Using Eqs. (8.21) and (8.22) it is easily shown that A, is of the form
(8.25)
(8.26)
This expression represents the rate of dissipation per unit initial volume
and is positive. Note that Eq. (8.26) is a consequence of relation (8.2.5),
which is more general and concerns a virtual change, whereas the actual
change in Eq. (8.26) is a particular case.
When in addition to being linear the relations between the rates and the
dissipative forces satisfy Onsager’s reciprocity relations (1930, 193 I),
they may be written
C!J$= aCP/a&_ 3;” = $JrM/an;l” I 9 %TT = d91TM18Si, (8.27)
where ‘Sch is a quadratic form in [, and GJTMa quadratic form in i# and
Si, with coefficients functions of the local state.
Applying Euler’s theorem, we write Eq. (8.26) as
Using Eqs. (8.13) and (8.14) and the definition (7.15) of !+:, this yields
with
From Eqs. (8.14), we may obtain fii in terms of X,! and h,“. Substituting
these values of rii into CR”determines the latter in terms of local state
variables, their gradients and i@. Thus in the quantity
&k = ~~“c~;) (8.35)
the only fluence fields are the state variables Mj" . With gi(Xj, xj’) repre-
senting a fluence field due to a unit concentrated source at x,?, Eq. (8.33)
may be written
6Hi = 8%: - J$ Ek SMF + J,- gi(X,, ~7) TJ;” 6Mj da+, (8.36)
New Variational-Lagrangian Irreversible Thermodynamics 29
(8.38)
Thus 6Si and Si are now expressed in terms of the fluence vectors S,? and
M!, which are state variables.
An important simplification occurs in problems where the effects of
inertia and body forces are negligible, and more generally ifjjb is negligi-
ble. In that case, withAk = 0, Eq. (8.33) yields
(8.39)
where
or
In many problems where fi” remains small these simplified results are
applicable.
Another possibility is to introduce the pseudo-entropy fluence S: de-
fined by Eq. (7.15). Equation (8.29) becomes
(8.44)
Hence
30 M. A. Biot
and
(8.47)
Thus 6Si and Si are now expressed in terms of state variables ST and ep.
Another important simplification also applies for problems where the
entropy produced does not contribute significantly to the total value of the
entropy Y. In that case we may write
Y = -dSf/dXi, (8.49)
and Si itself becomes a state variable. This is the case for quasi-reversible
processes and in linear thermodynamics.
where
(9.8)
.?k = (9.9)
32 M. A. Biot
It represents the work 6WM of the mechanical forces acting on the pri-
mary system plus the work 6 WTH accomplished in the thermobaric trans-
fer of heat and mass injected at the boundary. In addition, we generalize
d’Alembert’s principle by including in the external work the virtual work
-CiZi 6qi of the reversed inertial forces due to mass accelerations in the
primary system at any particular instant of the evolution. The generaliza-
tion involved here implies the validity of d’Alembert’s principle where U
is a thermodynamic energy involving heat and not simply a mechanical
potential as in classical mechanics. Accordingly, we may write
(9.15)
where
awrn = - I,ic $k 6MI + BIT 6Hi)ni dA (9.16)
6WGTH = -
JA(T pk 6Mf + 8 SSi)ni dA (9.20)
now includes the work against the gravity field in the boundary thermo-
baric transfer. We have put
‘Pk = $k + %, (9.21)
where % is the gravity potential field per unit mass chosen so that the
supply cells are on the surface % = 0. We have called $0kthe mixed
convective potential.
With time derivatives instead of variations, Eq. (9.14) yields the total
rate of dissipation as
T,$* = To I, S* da (9.22)
where itp At is the heat produced by the irreversibility in the time interval
At. If we have a thermal well at a lower temperature T,, this heat is not
entirely lost, because we may recover the mechanical work
For this reason we have called TP the intrinsic dissipation, and ToS* the
relative dissipation (Biot, 1975, 1976b).
In the modified thermobaric transfer heat is pumped from the thermal
well and injected into the various elements and cells of the hypersystem.
The process involves only mechanical work (given by Eq. 9.1) and pro-
vides purely mechanical definitions of qk as well as the exergy and 6 WTH.
The pump may use matter, subject to a Carnot cycle. However, it is of
interest to point out (Biot, 1976a) that use of the Carnot cycle may be
avoided by using pure heat as blackbody radiation extracted from the
thermal well and compressed adiabatically to the required temperature of
injection.
New Variational-Lagrangian Irreversible Thermodynamics 35
(10.1)
(10.2)
Substitution of these values into Eq. (9.19), which expresses the principle
of virtual dissipation, considering that 6qi is arbitrary, yields
Zi + 69’laqi + Ri = Qi, (10.4)
where Zi are generalized inertia forces, Ri are generalized dissipative
forces, and Qi are driving forces of a mixed mechanical and thermody-
namic nature due to environmental conditions (Biot, 1975, 1976b). These
equations are the general Lagrangian equations of evolution of irrevers-
ible thermodynamics. For systems that are quasi-reversible such that
local states do not deviate much from a local equilibrium satisfying Onsa-
ger’s (1930, 193 1) principle, the total virtual dissipation may be expressed
in terms of a dissipation function. For example, when the entropy produc-
tion is due to thermomolecular diffusion and chemical reactions near local
equilibrium, Eqs. (8.25), (8.27), and (9.14) yield
(10.5)
D=/$%!Cl=;.S* (10.7)
(10.9)
(11.1)
(11.3)
where p is the density of the mixture. The viscous stresses are then
written
oij = ql(uij + Vji) + rj2 aijV//, Vij = dVi/aXj, (11.4)
where r), and r/2 are viscosity coefficients depending on the local state of
the mixture. These stresses may be written in the form
‘+ij = lYSP/avij, (11.5)
where
38 M. A. Biot
(11.6)
(11.7)
We shall assume that the mass and heat flux obey locally linear laws
and Onsager’s principle. In that case the virtual entropy production due to
thermomolecular diffusion as derived from Eqs. (8.26) and (8.27) is given
by
T aS*TM = (11.10)
where % is the gravity potential field and Yf the exergy per unit volume.
The fixed-coordinate system defines cells whose volume is not varied.
Hence from Eqs. (3.22) and (9.6) with Sv = 0, we obtain
@j”= \@
P
A, atp+ T (dk 6Mk + 8 SLY) dR, (11.14)
I,(ai~~M~-CAp65,+C~k6Mk-B~6Si+T6s*)d~=0. (11.17)
P I
We then introduce the value (11.11) of T 6s” and the value (11.2) for Mk.
After integration by parts we equate to zero the coefficients of the arbi-
trary variations and obtain
(11.18)
aTlax.I + XZJTM/&$,
I = 0> (11.19)
-A, + CRp= 0. (11.20)
These are the field dynamical equations of the viscous fluid mixture with
reaction-diffusion. They bring out a fundamental coupling between diffu-
sion and the viscous stress gradient aa,laxj. They contain the unknowns
tPb, Mf, and Si. According to Eqs. (8.31) and (11.2), they determine the
state of the system as a function of time, if we may neglect the contribu-
tion of the entropy produced s* to the value Y of the entropy. However, if
this is not the case, we may determine s* by adding the auxiliary equation
which expresses the rate of dissipation and is obtained from Eq. (11.11)
by replacing the variations by time derivatives.
40 M. A. Biot
+
f
n+ gi(X[,
(
Xj+) T jjk SMjk+ Uij $) J
Lift+ (11.22)
+ (11.23)
(11.24)
where
(11.24a)
k
represents the total energy flux. Note that in this expression the quantity
& n;ri”E, + tii is the diffusive energy flux (8.41).
We have not yet mentioned the important fact that the dissipation func-
tion QTM for thermomolecular diffusion must be invariant under transla-
tion. It was shown (Biot, 1982a) that this condition is satisfied if we put
where
u!k
I
= 2l;i!/m,
I
- &k/m
I k, UkS
I = tiftmk - SiIY’y (11.26)
New Variational-Lagrangian Irreversible Thermodynamics 41
and Yc is equal to the value (4.17). The coefficients Cik and Ck are func-
tions of the local state, and the values of Ck are chosen such that the
coefficient of $ is T/2k where k is the local thermal conductivity of the
mixture. The dissipation function (11.25) satisfies the identity
ck (11.27)
Using this relation, we may verify that the field equations (11.18) and
(11.19) satisfy the total momentum balance (Biot, 1982a) by multiplying
Eq. (11.18) by mk and Eq. (11.19) by Yc. Adding the results and taking
into account the identity (11.27) and the modified Gibbs-Duhem theorem
(4.16) for v = 1, we obtain
Jnaj2 k
SMjkdfl = ~Z;6q;. (11.32)
(11.33)
(11.36)
is the chemical dissipative force. It can be shown that the viscous dissipa-
tive force is
(11.37)
where $9’ has the value in Eq. (11.6). The dissipation function Dv is a
quadratic form in 4;. The value, of %,rM is the dissipative force due to
thermomolecular diffusion. It is obtained by writing the variational rela-
tion
replacing 6Si and si by the values obtained from Eqs. (8.37) and (8.38).
This provides the exact value of RTM as a nonlinear function of Gi. How-
ever in most problems we may neglect the integral in the values (8.37) and
(8.38). In this case we may write
6lDTM
RTM
I
= - DTM = -rbc#M &
(11.38a)
&ji ’ I
s1 Tt
~Q~6qi=6W”-IA(~~xSM:+B6Si)nidA, (11.39)
where A is the boundary of R. On the right is the virtual work of all the
mechanical forces. The integral represents the work due to thermobaric
transfer at the boundary considered as fixed, whereas 6WM represents the
virtual work of all other forces. Special care must be exercised here by
noting that in the absence of viscous stresses 6WM = 0. In this case the
virtual work due to boundary displacement 6ui against the local equilib-
rium pressure p is already included in cpkSM/ of the surface integral.
Hence the remaining work 6WM is due only to the viscous stresses and is
written
(11.42)
as
where
and
44 M. A. Biot
1
SBi =JjW;j, Wij = 3 (Uij - Uji). (11.45)
where Q-:is the elastic part of the stress for reversible deformations and C,
is the heat capacity of the solid per unit initial volume with constant
strain. The value of h$ is given by Eq. (6.7) without calorimetric measure-
ments as
h$ = - T(d$dT), , (12.4)
New Variational-Lagrangian Irreversible Thermodynamics 4.5
where
r; = T&, T, x,) (12.5)
represents the local equations of state of the solid with the elastic stresses
expressed in terms of strain and temperature. The subscript E in Eq. (12.4)
indicates constant strain. Since the solid may be intrinsically nonhomoge-
neous, r: may depend also on the initial coordinates Xi.
Integration of Eqs. (12.2) and (12.3) yields %, Y, and ‘V = Q - TOY as
functions of Eij, T and xi.
The entropy production arises from the viscous stresses and the ther-
mal conduction. This provides another example of two distinct types of
entropy production, the first being due to an irreversible production of
heat while the second is not.
In order to apply the principle of virtual dissipation we must evaluate
the virtual entropy production. We therefore replace the differentials by
variations in Eqs. (12.2) and (12.3). They yield
6% = r; 8~ + T 69. (12.6)
For an irreversible transformation, conservation of energy is expressed
by
6% = (7; + rz) a&ij - d 6HildXi, (12.7)
where ri is the additional viscous stress and 6Hi is the variation of con-
travariant heat fluence.
We recall that fii is the heat flux across a deforming material area
initially normal to xi and equal to unity, whereas Hi is its time integral.
Equating (12.6) and (12.7) yields
69 = 6s” - d GSildXi 3 (12.8)
where
T 6s” = TV
u &s..
rJ - aTlax.I 6s. I? 65‘; = 6HiIT. (12.9)
We recognize the term (-dT/axi) 6Si already derived in (8.6) for thermal
diffusion. The first term may be interpreted as due to heat production
(Biot, 1981, 1982b) by considering an adiabatic transformation where con-
servation of energy requires
&?L = (7: + r$) SEij. (12.10)
Obviously this is equivalent to a reversible transformation where an
amount of heat 6hP = TzSEij is added in order to reproduce the same
change of state as in the irreversible process. The heat corresponds to the
concept of uncompensated heat of Clausius and illustrates the physical
46 M. A. Biot
difference between the two types of entropy production in Eq. (12.9). The
same distinction was discussed above in Section VIII.
As in the case of thermomolecular diffusion (see Section VIII) we now
introduce the kinetics of irreversibility. We write
(12.16)
where the volume integral is for the initial domain CRof the space Xi and p
is the initial density. The variation of 9 is
8~=6~-T”69=7~6e,+86Y (12.18)
(fJ = T - To). The variations are applied only inside the domain a, and
with the values (12.16), (12.17), and (12.18), the variational principle
(9.19) is written
= a.9,,laa,j
8&/.&V 6aij, 6aij = d8uildXj. (12.22)
After integration by parts, we equate to zero the coefficients of the arbi-
trary variations 6Ui, 6Si and obtain
(12.23)
g + TAySj = 0. (12.24)
I
These are the field equations for the deformable solid with elastic and
nonlinear viscous stresses and with thermal conduction.
In solving the problem we may consider ui and St as unknowns that
determine the state through Eq. (8.30) if we also determine the entropy
produced s*. This may be done by adding to the field equations the auxil-
iary equation
TS” = ~~6~ + TAuSiSj (12.25)
derived from Eq. (12.15). It extends Meixner’s (1941) result to deformable
solids with viscosity. Another procedure in analogy with the use of Eq.
(8.36) for thermomolecular diffusion is to introduce a pseudoenergy
fluence 9: defined by Eq. (7.15). The energy balance equation (12.7) is
then written
-di!@:laXi = ($ + T;) 8~ - dMIi/aXi. (12.26)
Hence
48 M. A. Biot
6H; = T 6Si = 8%: + lfl+ g(x/ 9 x:)(7; + ~z) 8~0 do+. (12.27)
This becomes
g+z=o.
I I
(12.33)
where fij is the Cartesian heat flux in the Xi coordinates. By using Eqs.
(7.12) we introduce the contravariant heat flux and obtain
(12.37)
(12.38)
showing the covariant nature of A,. Note that A,, is a function only of the
strain Ed, T, and Xi, and hence represents an intrinsic physical property,
whereas A, is referred to the fixed axes xi and depends also on the rotation
of the material.
Lagrangian equations are immediately obtained from the principle of
virtual dissipation. For simplicity, we shall assume that the entropy pro-
duced s* does not contribute significantly to the state variables, so that
the system may be described in terms of generalized coordinates qi by the
displacement field Ui and the contravariant entropy fluence Si. We write
where
+’ n ptiiicidC! (12.42)
I
2
(12.44)
50 M. A. Biot
where
(12.45)
(12.47)
where
(12.50)
With 69(qJ and the values (12.41), (12.44), and (12.50), the principle of
virtual dissipation (9.19) leads to the Lagrangian equations
(12.51)
(12.53)
(12.54)
of the temperature T and the Jacobian J (7.1 l), which represents the ratio
of final to initial volume of each fluid element. If the fluid is nonhomoge-
neous, %. and Y are also functions of the initial coordinates xi.
The thermal dissipation function per unit initial volume is derived from
Eq. (12.30) as follows. The thermal conductivity
The thermal dissipation function per unit initial volume (12.30) is there-
fore
(13.5)
(13.6)
where
71 = ql(J, T, -4, ~2 = ~(59 T, xi)
are viscosity coefficients, functions of the local state variables J, T and
the initial coordinates Xi. If there is no bulk viscosity we put
2
3 7)l + r/2 = 0. (13.7)
New Variational-Lagrangian Irreversible Thermodynamics 53
The virtual dissipation due to the viscosity per unit initial volume is
T 6~“” = Juip %&la?, = Jai+ dxjlaZ, 6aij (13.8)
where
6ati = d&.dJdXj. (13.9)
where
&ij = dlii/dXj. (13.14)
We substitute the value (13.13) into expression (13.6) and changej into p,
thus obtaining
rip = (~llJ)(Cpv& + C&p) + (s2/J)&~Cd/v. (13.15)
With this value, the Piola stress (13.12) becomes
Ti = (~~IJ)(C,“C~jki, + CpjC’,kp,) + (~2/J)GipC~jC’~,&~,. (13.16)
This may be written as
Ti = (~lIJ)(C~vC&~+ C+$iu)bpv+ (~lJ)C,Cpv~pv. (13.17)
By putting
By = (~llJ)(C/uC$i~ + CpjCiv>
+ (~/J)c~cpv, (13.18)
54 44. A. Biot
.. d a%
Wi - G, <TF+ Tz) + p axi = 0, (13.27)
where % is the gravity potential, %I*is given by Eq. (13.5), and p is the
initial density. An auxiliary equation for S* is
New Variational-Lagrangian Irreversible Thermodynamics 55
(13.31)
and F1, F2, F3 are functions of the three invariants e&, ekei, e;ejke;i.
In the general case they are also functions of T and xi if the fluid is
nonisothermal and nonhomogeneous. A very simple proof of this formula
has been given (Biot, 1976b). The Piola viscous stress TT in this case
cannot generally be derived from a dissipation function as for the Newto-
nian case (13.21). However, it is given by the same formula (13.12) in
terms of the stress Q. The field equations are then the same as Eqs.
(13.27) and (13.28). However, the Lagrangian equations become
(13.32)
where
(13.33)
(13.35)
(13.37)
(14.2)
(14.3)
New Variational-Lagrangian Irreversible Thermodynamics 57
The thermal time history of the domain is determined by the fluence field
Hi in terms of a finite number of generalized coordinates whose evolution
is governed by the Lagrangian equations
aV/dqi + aD/aGi = Qi. (14.4)
The thermal generalized force Qi is obtained as a particular case of Eq.
(12.50) by putting 6Ui = 0 SO that
(14.5)
which is the variational principle derived earlier (see Biot, 1970). It leads
to the Lagrangian equations (14.4) with a thermal potential V, a dissipa-
58 M. A. Biot
tion function D, and a generalized thermal force Qj, which are now ex-
pressed as
(14.9)
This result may be derived directly (Biot, 1970) or also as a particular case
of the general treatment of Section XI when restricted to pure heat trans-
fer for a given velocity field of the fluid.
Attention should be called to special formulations such as that of asso-
ciatedJluencefields, which lead to the use of scalar temperature fields as
unknowns instead of Hi, and the treatment of boundary heat transfer to a
moving fluid using the concept of a trailing function (Biot, 1970). The
latter eliminated the inconsistencies of standard methods based on local
heat transfer coefficients. Collective analysis by Lagrangian equations is
ideally suited to the unified treatment of heat transfer in mixed systems
constituted by solids and moving fluids.
A highly useful concept that was derived from the Lagrangian formula-
tion is that of penetration depth, which yields immediately the heat
fluence due to sudden temperature rise at the boundary. It can be used as
a basic tool simplifying the formulation of very complex problems.
Application of the Lagrangian equations has brought simplification and
physical insight in many problems of heat transfer (Lardner, 1963, 1967;
Prasad and Agrawal, 1972, 1974; Chung and Yeh, 1975; Yeh and Chung,
1977).
Many types of finite element methods may also be derived directly from
the Lagrangian equations. For example, we may treat as generalized co-
ordinates the fluence vectors located at the vertices of a grid, using linear
or polynomial interpolations of these values to represent the complete
New Variational-Lagrangian Irreversible Thermodynamics 59
Eq. (15.2) we may obtain T as a function of Y, Q, 5,) M’, and xi. Thus the
exergy
=V= Q_l- T&f’ (15.3)
per unit initial volume becomes a function of these variables. In a gravity
field %(Xi) the mixed collective potential of a solid occupying the initial
domain R is
dMl‘
y=Jp+( mo + m)%(Lf;)] da, m=-CL
k axj ’
(15.4)
where mo is the initial mass per unit initial volume and m is the mass
added by the contravariant fluences Mf. We recall that kl: is the mass flux
of the dissolved substance k through a deforming solid area initially equal
to unity and initially normal to the xi axis. From Eqs. (6.11) and (15.3) we
obtain
6~=rij~~ij+fC~8Mk+OSY--AAp~~p. (15.5)
k P
This result extends the value (12.18) to the solid with molecular diffusion
and chemical reactions.
The virtual dissipation is due only to the chemical reactions and the
thermomolecular diffusion. It has a form similar to Eq. (8.25).
(15.7)
complexity of the physics. We may consider 5,) Ui, MF, and Si as the field
unknowns, describing the evolution of the state of the system. However if
the contribution from s* is significant in determining the state of the
system we may add an auxiliary equation such as Eq. (11.21) for the rate
of dissipation or proceed as in Sections XI and XII by introducing a
pseudo-energy fluence St in the field equations.
By neglecting s* in first approximation we may describe the system in
terms of generalized coordinates qi as
(15.16)
DTM ;/$3TMdfi
= _ (15.17)
is due to forces fi applied at the boundary per unit initial area and to
thermodynamic forces of the environment on the open system. This ex-
pression contains the particular case [Eq. (12.50)] derived above.
For the case where chemical reactions are not far from equilibrium, the
evolution is quasi-reversible, and we may introduce a chemical dissipa-
tion function
(15.19)
62 M. A. Biot
(16.9)
The principle of virtual dissipation has been used to develop the field
and Lagrangian equations for the dynamics of a porous solid with fluid
saturation of the pores and heat transfer by convection and conduction.
We shall follow the procedures developed earlier for the isothermal non-
inertial case (Biot, 1972) and later for the nonisothermal dynamic case
(Biot, 1977b).
For the noninertial case the problem is similar to that of a solid with a
single substance in solution, and in many ways it may be considered as a
particular case of the one treated in Section XV. However in the present
New Variational-Lagrangian Irreversible Thermodynamics 65
case we shall also take into account the inertial forces due to the motion of
the porous frame and the motion of the fluid relative to the frame. We
start with a unit initial volume of the material, initially of total mass mo.
As the material is deformed, the exergy per unit initial volume is
Y = Y(E~, m, T, XI), (17.1)
where cij are the six strain components as defined in Section V, and
m = -aMilaXi (17.2)
is the mass of fluid added in the pores per unit initial volume. The con-
travariant fluence Mj is the mass of fluid that has been flowing through a
material surface of the porous frame initially equal to unity and initially
normal to Xi. A material point of the solid frame initially of coordinates Xi
is displaced to a point of coordinates Xi = xi + Ui. If the porous material,
which is approximated as a continuum, is inhomogeneous, Y is also a
function of the initial coordinates xi. Using procedures similar to those in
Section XV, we may evaluate the entropy Y of the element in terms of .Q,
m, T, and xI. Hence we express T replacing T by Y as one of the state
variables. We write
‘Y = Y(eij, m, 9, XI), (17.3)
with the property analogous to Eq. (15.5):
(17.4)
where 8 = T - To. The fluid pressure pf in the pores in the present case is
equal to the injection pressure, which was denoted by ps for the case of a
substance in solution. The corresponding injection enthalpy Fr and injec-
tion entropy Ff are given by Eqs. (2.2) and (2.4) as
(17.5)
where pf , pE T’, and F; are the pressure, density temperature, and specific
entropy of the fluid along the path of thermobaric transfer from a supply
cell of fluid at the pressure p. and temperature To. The convective poten-
tial is
d, = Fr - Ts;-, (17.6)
and the mixed collective potential is now
(17.9)
(17.12)
Hence
Ati = (mo + RZ)tii + tij CY.iT~laXj. (17.13)
The virtual work of the inertial forces for the porous solid, when we vary
6ui and 6Mi, is
New Variational-Lagrangian Irreversible Thermodynamics 67
C, Zi 69,
1
= I,,9i SUMda + lE iii 6M/ dY, dX2 dZ, 3 (17.15)
By varying Sui and 6Mi inside the domain fi and applying the principle
of virtual dissipation, we proceed as in Sections XII and XV. This leads to
the field equations
(17.18)
+ ii (17.22)
(17.23)
(17.24)
The forces fj are applied at the boundary per unit initial area.
A semilinear theory of deformation of porous solids was also developed
for the particular case where linear dependence is assumed between fluid
pressure and microscopic volume changes while bulk deformations re-
main nonlinear (Biot, 1973~).
The problem of fluid flow through a rigid porous solid with heat transfer
was also discussed in detail (Biot, 1978), including phase changes from
vapor to liquid and multiple diffusion channels due to surface adsorption.
The results lead to a large variety of possible applications in geothermal
and aquifer problems with heat and fluid flow including associated sub-
sidence.
where h is the heat acquired per unit volume and Fokis the initial value of
the injection entropy in the initial equilibrium state. Since Y and Mk are
state variables, the heat injected h is also a state variable.
Another simplification, which is a consequence of the first, is the use of
the thermobaric potential $k instead of the convective potential pk. For
example, we may write Eq. (15.5) for the more general case of a solid, in
differential form, as
h = -aHi/E+Xi (18.3)
(18.5)
dY = dhlT. (18.9)
By using the equations of state (18.7) in (6.7), the value (6.4) of dh be-
comes
dh = T& deij + c do, (18.10)
where c is the heat capacity per unit volume at constant strain, which we
shall assume constant. We integrate the values (18.8) and (18.9) along a
path, first for 8 = 0 and then for deij = 0. Keeping only first- and second-
order terms, we derive
Q = $CcL”&..& + T,,sT, (18.11)
rJ u P
9 = sT - $(ce*/T;), (18.12)
with
70 M. A. Biot
We see that sT is a first-order state variable, but it does not represent the
correct entropy to the second order. The thermoelastic potential or ex-
ergy becomes (Biot, 1981)
‘V = ‘Q - TOY = B(C;V~ti~pv+ cd*/T,,). (18.14)
(18.15)
where uji(xl) are fixed displacement fields. Fluence fields Mj” and Sj as well
as <,, are represented similarly. Application of the principle of virtual
dissipation then yields the linear Lagrangian equations for the perturba-
tions qi. They are
(18.16)
(18.17)
a dissipation function
(18.18)
(18.19)
Qi = C Zvqj> (18.23)
j
with
Z(j=Zji=CDg&+Dij+D~p,
s s
df
&f(t) = em” s’e H’ -$ dt’,
0
PfW = z. (18.25)
(18.27)
where the operators ar, i&, and & are of the type (18.27). Note that the
heredity properties are due not only to viscoelastitiity of the solid itself
but also involve interactions of a very genera1 relaxation type between the
fluid and the solid, such as fluid squirting in microcracks and between
grains, microthermoelasticity, mutual solubility, adsorption, and surface
diffusion as well as chemical reactions.
A principle of viscoelastic correspondence is also valid for porous me-
dia whereby all formulas of the purely elastic theory are valid for the
viscoelastic case if one simply replaces the elastic’coefficients by the
corresponding operators of the stress-strain relations (18.28) and (18.29)
(Biot, 1962).
The linear thermodynamic theory has also been applied to piezoelectric
crystals with thermal dissipation (Mindlin, 1961, 1974) by adding suitable
electrostatic terms.
New Variational-Lagrangian Irreversible Thermodynamics 73
(19.4)
74 M. A. Biot
where vij is a second-order quantity in au. For example, in the case (5.5),
we obtain
r)ij = t(eipopj + &jfiWpi + Wp6Jjp)y (19.5)
or
where
The exergy V is a function of Q, 5,) Mk, and sT. Because S, dcij + & qbOk
dMk is an exact differential, dW is also an exact differential. Furthermore
we may consider au as independent state variables. In a linear theory we
need evaluate @V only to the second order, hence we may replace Ed by eij
in the value of dW. We write
G=I,F
mo%iui + (%o + %iUi)C Mk + i mo%ouiuj dfk.
k 1 (19.16)
We now apply the principle of virtual dissipation (9.19). With the value
(19.8) of dY it becomes
2 Zi 6qi + 69 + To 6S*
I
where
+ 2
k
m&iii + af) 6&
1
da, (19.21)
where mokis the initial mass of substance k per unit volume of solid and a:
its acceleration relative to the solid. We denote by SUP the virtual dis-
placement of the substance associated with 6M/. We have in the linear
case
a? = ti//mok, 6U; = #&%?‘lok. (19.22)
With these values, (19.21) may be written in the form
(19.23)
where
(19.24)
where
9 = Gj@ + QrM (19.26)
The first term Sch is a quadratic form in [,, and represents the dissipation
due to chemical reactions. The second term represents the dissipation due
New Variational-Lagrangian Irreversible Thermodynamics 77
(19.27)
aTlaxi + aWa,$T
= 0, (19.29)
-A, + aWa& = 0, (19.30)
where
AKi= mo%guj
+ %iC Mk, (19.31)
k
(19.33)
where
(19.34)
D = I, 9 dQ (19.35)
(19.37)
(20.1)
or, explicitly,
(20.2)
Solutions of these equations are of the type exp(pt) where the p are
characteristic roots of the system. The roots p are either real or complex
conjugate and the system is stable if the real part of each root is negative.
However if there are roots with positive real parts the system is unstable.
A fundamental theorem has been established that states that unstable
roots are all real, hence that the instability is always nonoscillatory (Biot,
1965b, 1974). To show this, we consider a root p and its complex conju-
gate p*. The roots satisfy the equations
We multiply Eqs. (20.3) by qT and Eqs. (20.4) by qi and add the results.
Taking into account the symmetry properties mij = mji, b, = bji , and aij =
aji, we obtain
P + P* ’ 0, (20.9)
and Eq. (20.7) cannot be verified unless p = p*, i.e., unless the root is
real.
Thus near unstable equilibrium the perturbations in the linear range are
nonoscillatory and proportional to increasing exponentials. Note that this
includes dynamical systems with inertial forces.
Another property is derived by assuming an unstable solution with real
p. Equation (20.5) shows that in this case we must have
9i = (oi(t> (23.2)
where Aqi is a small perturbation. The driving force Qi is given and main-
tained unperturbed. Substitution in Eqs. (23. l), neglecting higher-order
terms, yields
c- j
a29
Qi Nj
Aqj + ARi = 0. (23.4)
(23.7)
84 M. A. Biot
C bi AGj = 0. (23.9)
j
By multiplying these equations by Aqi and adding the results, we obtain
then Eq. (23.13) shows that AD must decrease with time. Because it is
positive definite, the values of Aqi must also tend to zero. Hence the
inequality (23.14) constitutes a fundamental stability criterion.
If the inequality (23.14) is not verified, instability may arise. It may be
due to the negative nature of 8*9/dqi aqj, in which case it is analogous to
the instability of a linear system near equilibrium considered in Section
20. The instability may also be due to the negative nature of (dR;/dqj +
dRj/dqi), which arises essentially from the nonequilibrium state of evolu-
tion. Such a case of instability was illustrated by the example of an em-
bedded viscous layer (Biot, 1976b), as recalled in more detail in the next
section. In addition, in this example & = &ji, which implies that the
instability is nonoscillatory.
New Variational-Lagrangian Irreversible Thermodynamics 85
FIG. 2. Stack of alternate layers of large and small viscosity between rigid planes subject
to a compression P. Note the appearance of internal folding of wavelength L.
c Qpt = 0
s
and
(25.3)
REFERENCES
tion and Flow of Solids” (IUTAM Colloquium, Madrid, 1955), pp. 251-263. Springer,
Berlin.
Biot, M. A. (1956b). Thermoelasticity and irreversible thermodynamics. J. Appt. Phys. 27,
240-253.
Biot, M. A. (1958). Linear thermodynamics, and the mechanics of solids, Proc. Third U.S.
Nut. Gong. Appl. Mech., pp. l-18. ASME, New York.
Biot, M. A. (1962). Generalized theory of acoustic propagation in porous dissipative media.
J. Acoust. Sot. Am., 34, 1254-1264.
Biot, M. A. (1963). Theory of stability and consolidation of a porous medium under initial
stress. J. Math Mech. 12, 521-542.
Biot, M. A. (1964). Theory of buckling of a porous slab and its thermoelastic analogy. Trans.
ASME, Ser. E 31, 194-198.
Biot, M. A. (1965a). Further development of the theory of internal buckling of multilayers.
Bull. Geol. Sot. Am. 76, 833-840.
Biot, M. A. (1965b). “Mechanics of Incremental Deformations.” Wiley, New York.
Biot, M. A. (1966). Three-dimensional gravity instability derived from two-dimensional
solutions. Geophysics 31, 153-166.
Biot, M. A. (1967). Rheological stability with couple stresses and its application to geologi-
cal folding. Proc. R. Sot. Ser. A, 298, 402-423.
Biot, M. A. (1970). “Variational principle in heat transfer.” Oxford Press.
Biot, M. A. (1972). Theory of finite deformations of porous solids, Indiana Univ. Math. J.
21, 597-620.
Biot, M. A. (1973a). Nonlinear thermoelasticity, irreversible thermodynamics and elastic
instability. Zndiana Univ. Math. J. 23, 310-335.
Biot, M. A. (1973b). Buckling and dynamics of multilayered and laminated plates under
initial stress. Znt. J. Solids Struct. 10, 419-451.
Biot, M. A. (1973~). Nonlinear and semilinear rheology of porous solids. J. Geophys. Res.
78, 4924-4937.
Biot, M. A. (1974). Thermoelastic buckling. An unstable thermodynamic equilibrium at
minimum entropy. Bull. Cl. Sci. Acad. R. Belg. 60, 116-140.
Biot, M. A. (1975). A virtual dissipation principle and Lagrangian equations in nonlinear
irreversible thermodynamics. Bull. Cl. Sci. Acad. R. Belg. 61, 6-30.
Biot, M. A. (1976a). New chemical thermodynamics of open systems. Thermobaric poten-
tial, a new concept, Bull. Cl. Sci. Acad, R. Belg. 62, 239-258; Erratum 62, 678.
Biot, M. A. (197613).Variational-Lagrangian irreversible thermodynamics of nonlinear ther-
morheology. Q. Appl. Math. 34, 213-248.
Biot, M. A. (1977a). New fundamental concepts and results in thermodynamics with chemi-
cal applications, Chem. Phys. 22, 183-198.
Biot, M. A. (1977b). Variational-Lagrangian thermodynamics of nonisothermal finite strain
mechanics of porous solids and thermomolecular diffusion. Znt. J. Solids Struct. 13,
579-597.
Biot, M. A. (1977~). Variational-Lagrangian irreversible thermodynamics of initially
stressed solids with thermomolecular diffusion and chemical reactions. J. Mech. Phys.
Solids 25, 289-307; Errata 26, 79.
Biot, M. A. (1978). Variational irreversible thermodynamics of heat and mass transfer in
porous solids: new concepts and methods. Q. Appl. Math. 36, 19-38.
Biot, M. A. (1979). New variational-Lagrangian thermodynamics of viscous fluid mixtures
with thermomolecular diffusion, Proc. R. Sot. London, Ser. A. 365, 467-494.
Biot, M. A. (1981). Generalized Lagrangian thermodynamics of thermorheology J. Thermal
Stresses 4, 293-320.
New Variational-Lagrangian Irreversible Thermodynamics 91