Introduction To The Thermodynamics - Ericksen JL
Introduction To The Thermodynamics - Ericksen JL
Introduction To The Thermodynamics - Ericksen JL
J.L. Ericksen 5378 Buckskin Bob Drive Florence, OR 97439-8320 USA Editors
J.E. Marsden Control and Dynamical Systems, 116-81 California Institute of
Technology Pasadena, CA 91125 USA L. Sirovich Division of Applied
Mathematics Brown University Providence, RI 02912 USA To my dear wife and
friend, Marion E. Ericksen Mathematics Subject Classification (1991): 73B30,
7301,76A 15, 80A20 Library of Congress Cataloging-in-Publication Data
Ericksen, J. L. (Jerald L.), 1924- Introduction to the thermodynamics of solids /
Jerry L. Ericksen. -- Rev. ed. p. cm. -- (Applied mathematical sciences; 131)
Includes bibliographical references and index. ISBN 0-387-98364-3 (hardcover:
alk. paper) 1. Materials. 2. Thermodynamics. I. Title. II. Series: Applied
mathematical sciences (Springer-Verlag New York Inc.); v. 131. TA403.6.E74
1997 621.402' 1c21 97-37938 Printed on acid-free paper. The fin-st edition of
this book was published by Chapman & Hall, UK, � 1991. � 1998 Springer-
Verlag New York, Inc. All rights reserved. This work may not be translated or
copied in whole or in part without the written permission of the publisher
(Springer-Verlag New York, Inc., 175 Fifth Avenue, New York, NY 10010,
USA), except for brief excerpts in connection with reviews or scholarly analy-
sis. Use in connection with any form of information storage and retrieval,
electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed is forbidden. The use of general
descriptive names, trade names, trademarks, etc., in this publication, even if the
former are not especially identified, is not to be taken as a sign that such names,
as understood by the Trade Marks and Merchandise Marks Act, may accordingly
be used freely by anyone. Production managed by Steven Pisano; manufacturing
supervised by Jacqui Ashri. Photocomposed pages prepared from the author's
LaTeX files. Printed and bound by Maple-Vail Book Manufacturing Group,
York, PA. Printed in the United States of America. 9 8 7 6 5 4 3 2 1 (revised
edition) ISBN0-387-98364-3 Springer-Verlag New York Berlin Heidelberg
SPIN 10647553
viii Preface Preface ix this way, the course has helped to fill in the gap in
education mentioned at the beginning. Chapter 10 is part of the package for such
readers. This covers some of the difficulties encountered in trying to apply
thermody- namics to obtain a better understanding of the phenomena
encountered in solids, partly to indicate the need to grasp the basic concepts of
classical thermodynamics. Unfortunately, different experts in thermodynamics
have discordant ideas as to what these basic concepts are. In as elementary a
manner as possible, therefore, I discuss what they are as I understand them.
How- ever, this chapter is not so elementary as it deals with matters which are
unsettled and controversial. In this treatment, classical thermodynamics is
interpreted to exclude an important branch based on molecular theory, that is,
statistical thermodynamics, only because covering this in any reasonable way
would make the notes excessively long. The intent is to provide a small bridge to
newer work in thermodynamics. I have tried to choose a few references which
seem likely to be useful, rather than attempting to include all that may be of
interest. Since the notes have been used by readers from quite varied
backgrounds, references cited range from the very elementary to rather
sophisticated works. Readers will need to pick a subselection. I do not intend
this to be a replacement for other elementary books on thermodynamics and do
assrune that the reader has a little familiarity with the subject. For those who feel
a need for supplementary reading, I note a few of the many possibilities. I do
think it desirable for writers of elementary books to better cover applications to
solids. Historically, the subject first emerged from studies of ancient heat en-
gines, with a corresponding emphasis on dynamic processes. In particular, this
produced early ideas about energy and entropy as they are related to the old laws
of thermodynamics. Much in this spirit, although more mod- ern in style, is the
book by Truesdell and Bharatha [1]. What has become a more conventional view
is that energy and entropy are related more to equi- librium states although one
may be dealing with nonequilibrium processes. For a development of the subject
from this point of view, the reader is referred to the work by Kestin [2], for
example. He is influenced by Gibbs's ideas concerning equilibrium theory,
which will be discussed here. One of the most ardent proponents of the notion
that energy and entropy must be related to equilibrium states is Tisza [3],
although he is very critical of Gibbs's ideas. It is not clear how many really
believe that this view is a "law" of thermodynamics, but most writers of
elementary books abide by it. Included are some who exhibit independence of
thought in considering various other ideas about thermodynamics. In this
category is Pippard, whose book [4] is interesting for its comments and
examples, including some relating to solids. Others, myself included, see no
grounds for accepting the restriction and some reason to consider alternatives. In
the first nine chapters of this book, the theories treated do fit this mold fairly
well, with some caveats mentioned in Chapter 10. However, in Chapter 2, I
introduce ideas which are now being used frequently by those willing to consider
alternatives. This is one way of gaining experience with the Clausius-Duhem
inequality which is accepted and used in studies of irreversible processes by
some who, in practice, do accept the conventional view mentioned above, Kestin
in particular. Many elementary books do not mention this old inequality. Rather
obviously, it is probable that theories of equilibrium will fit the conventional
mold and it would have been easy to include many more examples of this kind.
For example, the basic ideas are made available to take the three-dimensional
linear theory of thermoelasticity and deduce all the useful inequalities satisfied
by moduli using the Clausius Duhem inequality and the thermodynamic theory
of stability. My experience is that this is not a matter of common knowledge
among experts in this area, but applications of this kind are fairly routine and
useful. It is also very easy to find theories of solids that are commonly used,
some quite old, which do not fit the conventional mold, as was mentioned
earlier. It can then be very difficult to know how best to try to apply ideas of
thermodynamics. One who has wrestled hard with problems of this kind is apt to
see the subject in a somewhat different light. In this category is Bridgman [5],
who had a strong interest in applications to plasticity, in particular, although he
had no great success in mastering the difficulties involved. Readers unfamiliar
with elementary continuum mechanics may find helpful the text by Bowen [6].
Particularly in discussions with chemists I have encountered another prej- udice.
Roughly, it is that if a material really attains equilibrium, any shear stress must
have relaxed to zero. Perhaps this is why so many authors of works on
thermodynamics consider only problems of solids subject solely to a hydrostatic
pressure, if at all. In this respect most of the examples to be considered illustrate
rather common practices of users which tend to be ignored by authors of texts.
Considering how the various structures we make tend to deteriorate, I do
concede that those chemists have a point. If we grant it, equilibrium theory for
solids should be similar to that for fluids. Then the different theories of
thermoelasticity which appear to be equilibrium theories are not really of this
kind, despite appearances. If I accepted this and Tisza's view, as I interpret it,
thermoelasticity theory is based on improper usage of thermodynamics. I think
that it may well be that there is something deeper to be understood here, which
could well influence and improve our understanding of energy and entropy. In
Chapter 10, I will say a little more about this. However, it often happens that
theories which prove to be successful were arrived at by infirm reasoning and
this alone is no reason to reject them. I subscribe to the view that, if we can
understand more fully why such theories succeed, we will improve our chances
of constructing a still better theory.
2 Constitutive Transfer for Theory of Heat Bars and Plates 2.1 Thermodynamics
of Rigid Bars In this section, we consider the one-dimensional theory of heat
transfer in rigid, stationary bars, illustrating how some of the general
thermodynamic ideas are interpreted in this context. Mathematically, the points
on a bar are represented by points in the interval 0 < x < L. (2.1.1) Quantities of
interest, such as temperature, will then, in any process, be functions of x and t.
We require a theory capable of describing the possible temperature distributions.
One rather general idea is that a part of a ther- modynamic system can itself be
considered as a thermodynamic system, so, in particular, any part of the bar
qualifies. In Chapter 8, we will note that it is not always feasible to use this idea.
Forces may act on any such part but these will do no work because motion is
excluded. Thus, any such part, or the whole bar, will be mechanically isolated, P:
0. (2.1.2) Another general idea is that energy and entropy are additive. 1 Here,
the traditional way of covering this is to assume that they are representable by
1Generally, thermodynamicists call such quantities extensive variables, refer-
ring to quantities which are not, for example O, as intensive variables.
12 2. Constitutive Theory of Heat Transfer for Bars and Plates integrals E--fedx,
$--frldx , (2.1.3) where thc integral can be over the interval (2.1.1), or a subset.
Physically, heat transfer can take place along the bar by conduction. To describe
this, we introduce a heat flux q. In terms of this Qc, the rate at which heat is
supplied by this mechanism to a subinterval (x,x2) is rcpresented by Qc --
q(x2,t) - q(x,t) = q x2. (2.1.4) With one-dimensional thcory, it is difficult to
describe transfer through the sides of the bar in a similar way, so we introduce
another representation, Qo -- rdx, (2.1.5) sometimes considered as a measure of
radiation, sometimes as a crude ap- proximation to the effects of conduction or
perhaps somc combination of both. Then, for the subinterval, (1.1.1) takes the
form dE d dt - dt e dx = Q = Qc + Q0 (2.1.6) = q(x2, t) - q(xl, t) q- r dx.
Assuming the functions are smooth enough, we can rewrite this as I dx ---- i XX
+ r dx, (2.1.7) and infer from the arbitrariness of x and x that Oe Oq - q- r.
(2.1.8) Ot Ox In this context, the accepted view is that the entropy is involved in
an inequality, the so-called ClausiusDuhern inequality, rldx _> (q/O) + (r/O)dx,
(2.1.9) reducing to the local form, Or> 0 _ xx(q/O ) + r/O. (2.1.10) 2.1
Thermodynamics of Rigid Bars 13 Additional equations are needed, and
physically these should take into account that the thermal response will be
different for different materials. Before considering this, let us consider some of
the different ways of describing the notion that the bar is in contact with a heat
bath at tem- perature 0B(t), a given function of t. One possibility, envisaging that
the side of the bar is coated with an insulator, is r = 0, O(O,t) = O(L,t) = OB(t).
(2.1.11) Then, applying (2.1.9) to the whole bar, we get dS d fo L dt -- dt rldx >
[q(L,t) - q(O,t)]/O, (2.1.12) agreeing with the Clausius Planck inequality (1.2.2).
Or, we might replace these temperature boundary conditions by boundary
conditions of the ra- diation type. With Newton's law of cooling, this would give
q(L, t) = a[O - O(L, t)], (2.1.13) q(0, t) = -ct[0 - 0(0, t)] with a a positive
constant. That a should be positive reflects the idea that heat should flow into the
bar if it is cooler than the heat bath, out if it is hotter. Then, it is easy to verify
that q(L, t)/O(L, t) = a[O - O(L, t)]/O > q(L, t)/OB (2.1.14) -q(O, t)/O(O, t) > -
q(O, t)/O. Putting this together with (2.1.4) and (2.1.9), again assuming r = O,
we get dS -- > Q/OB, (2.1.15) dt - also agreeing with (1.2.2). Another commonly
used radiation condition is the Stefan-Boltzmann law, which would give q(L, t)
=/[0 - 04(Z, t)], q(0, t) = -/[0 - 04(0, t)] (2.1.16) with a positive constant, again
leading to (2.1.15) when r = 0. One can consider heat transfer from the sides of
the bar while insulating the ends. Using, for example, Newton's law of cooling,
this gives q(O,t) = q(L,t) = 0, r = a[0 - O(x,t)]. (2.1.17) Again, ct should be
positive, to have heat flow into the bar when it is cooler and out of the bar when
it is hotter. Also, it follows that r(x, t)/O(x, t) _> r(x, t)/Os(t), (2.1.18)
14 2. Constitutive Theory of Heat Transfer for Bars and Plates and that (2.1.15)
again holds, now with Q = Qb- One can replace the in- sulated end conditions by
radiation boundary conditions, again confirming (2.1.15). So, under various
assumptions of this kind, the Clausius-Duhem inequality implies the Clausius-
Planck inequality. Unlike the latter, it also applies to cases where 02 might be
considered to depend on x, or to cases where 0 might be assigned different
values at the ends, in a bar with insu- lated sides. For classical studies of
equilibrium, one needs a constitutive equation depending on the material, and
one of the form e = e(V) (2.1.19) would fit Gibbs's preference. Let us consider
this bar in contact with a heat bath at temperature 0s = const. and explore the
consequences of definition III, in Section 1.3. We have L Es = 0 [e(V) - Ov]dx,
(2.1.20) defined for some functions (x). At let one of these should occur a
thermodynamic equilibrium state; call it . As possible variations of it, we
consider a one-parameter family of functions, as is commonly done in the
calculus of variations, where is the parameter and 5 an arbitrary smooth function
of x. A common idea is that any variation of this kind is possible, if it is small
enough, and, given 5, we can make it small by making small. For the moment,
consider 5 v fixed, and consider L E() = [e( + SV) - O,( + SV) ] dx. (2.1.21) For
small, we have, a first approximation, where Es(O) 5Es -- E(O) = (V) - Os 5vdx
(2.1.22) is called the first variation. We interpret the criterion for equilibrium as
meaning that we should have 6Es >_ 0, 2One of the better elementary treatments
of this subject is reference [11]. 2.1 Thermodynamics of Rigid Bars 15 and,
since/ can be chosen to be positive or negative, this requires that 5Es = O, for
arbitrary 5. By a fundamental lemma in the calculus of variations, this holds if
and only if d() = Os (2.1.23) Briefly, if this were not true, the bracketed quantity
in (2.1.22) would be positive or negative in some interval. Choosing 5 positive in
a subinterval of this and zero elsewhere produces a contradiction. So, (2.1.23)
gives an equation for determing (stable or unstable) equilibrium values of . With
other theories, similar reasoning is used in interpreting the criteria for equi-
librium. In equilibrium, one expects 0, the temperature of the bar, to match that
of the heat bath. This fits the common assumption that 0 and are always related
by de 0 = --. (2.1.24) d As a general proposition, it is certainly not safe to assume
that an equa- tion that holds in equilibrium also holds more generally. Here,
there is a consensus of opinion that the assumption has a range of validity which
is not restricted to equilibria but might well not include processes involving large
departures from equilibrium. Now, we can proceed to obtain stability conditions.
As a test for a local minimum, many would use the second derivative test, 52Es
= E(O) > 0 (2.1.25) or f0 2 dx 0 q which, by (2.1.24), leads to d d e . . (7)= k o.
(2.1.26) Briefly, if the second derivative were negative, it would be negative in
some interval by continuity and one could again find 5r/ for which the integral
inequality is violated. Conditions very similar to (2.1.23) and (2.1.26) emerge,
by similar rea- soning, in a great variety of theories. They represent what may be
viewed as minimal requirements for stable or metastable equilibrium. If (2.1.24)
gives 0 as a monotonically increasing function of r/, taking on all positive
values, then (2.1.23) holds for exactly one value of and (2.1.26) will be satisfied.
Some workers are happy to consider only constitutive equations
16 2. Constitutive Theory of Heat Transfer for Bars and Plates 2.2 Constitutive
Theory for Rigid Bars 17 having these properties. This excludes some
instabilities that might occur in principle, but I know of no clear evidence that
they occur in practice. According to the criterion, the most stablc equilibria
corrcspond to ab- solute minima. Clearly, the integral will be smallest if thc
integrand every- where takes on its minimum value? so the criterion for this is or
with (2.1.23), - oBv_> s(V) - oBv, de e(r/) - s(7) >_ -rr/(7)(r/- 7) D r/, (2.1.27)
where D is to be read as "for all." The aforementioned common assumption
implies that this holds for every choice of r/and 7. Mathematically, e is then a
convex function of r/. Were it not, (2.1.27) would require that (7, s(7)) be a point
of convexity of the graph of e = e(r/); the graph must be above the tangent line at
this point. For simplicity, make the assumption alluded to above, so we can solve
(2.1.24) to obtain r/as a function of 0, either quantity serving as a label for states.
The derivatives will satisfy d dO -- . dO d Note that, if equality holds in (2.1.26),
&l/dO will not be finite. Actually, this can happen at particular values of 0 at
which phase transitions take place, although some more general theory is needed
to describe what then occurs. 4 We will exclude this possibility. Now consider
starting with a bar in equilibrium at some temperature 0. Change this a little, to 0
+ A0z, and allow it to reach equilibrium. This will give a change in energy; by
(2.1.23) and (2.1.24) L AE= A o e(7) dx (2.1.28) LoXO ZXO ---- B ---- LOB B.
Here, the overbar denotes evaluation at the state corresponding to 7. Since no
work is done on the bar, this should also be the total amount of heat supplied to
the bar, which can be measured using a calorimetcr. Clearly L, 0 and A0B can
also be measured, so one can infer values of de/dO and drl/dO. By doing this for
a range of temperatures, one can get s and 3This excludes some commonly
ignored mathematical possibilities, like the function not being bounded below, or
not attaining a minimum. 4See, for example, the discussion of lambda transitions
by Pippard [4]. r/ as functions of 0, at least for some range of temperatures, and
from this, calculate e(r/). A careful worker will at least increase and decrease
temperature to see if thc results thus obtained are consistent. 2.2 Constitutive
Theory for Rigid Bars In the previous discussion, we said rather little about
some items, q in particular, but they play a more important role in the
consideration of heat transfer. The aim is to somehow relate these to
temperature. The classical theory of heat conduction uses Fourier's law q = n
O0/COx, (2.2.1) where n is a constant, or more realistically, a function of
temperature. This suggests that items of interest should depend on 0 and
COO/COx. Equilibrium studies suggest that e and r/ depend only on 0, but they
deal with cases where cOO/cOx could well be small enough to make a negligible
contribution. Thus, a likely assumption may be that we have constitutive
equations of the form s = s(0, CO0/COx), r I = rI(O, COO/COx), q: q(O,
COO/COx), (2.2.2) depending on the material from which the bar is made. The
usual idea is that these do not change when we put the bar in different
environments. In a slightly different category is r, which is not as independent of
the en- vironment from the examples of radiation laws. We do need some
prescrip- tion(s) for it, more or less like the examples discussed in Section 2.1.
The basic idea is to convert the energy equation (2.1.8) to a differential equa-
tion for 0 and to arrange that any solutions satisfy the Clausius Duhem inequality
(2.1.10). One way of proceeding is to eliminate r between thc two to obtain corl
x COt COq (2.2.3) O -0 ( q, O ) _> COt COx and to restrict the constitutive
equations (2.2.2), so this holds for essen- tially arbitrary smooth functions O(x,
t); one can respect obvious conditions such as 0 > 0. This procedure is not as
arbitrary as it may seem. For ex- ample, suppose that we accept Newton's law of
cooling together with the prescription for r given in (2.1.17) with the idea that 0
is an assignable constant. Choose the bar temperature function O(x, t) and think
of using (2.2.2) to calculate s(x, t), and so on. Then calculate COt COq f(x, t) --
COt COx' Now choose particular values of (x, t) = (x0, to), say, and solve f (xo,
to) = r(xo, to) = - O(xo, to)]
18 2. Constitutive Theory of Heat Transfer for Bars and Plates for 0B, to satisfy
(2.1.8) at (Xo,to). Clearly, (2.2.3) must then hold at (Xo, to). Actually, it is not
too unrealistic to assume that 0B is an assignable function of x, t. Accept this
and avoid evaluation at particular values of x and t. A weakness in the argument
is that the calculated values of 0 could be negative, and this is not physically
acceptable. Hence, we may sacrifice some generality in accepting the indicated
assumption. We now rearrange (2.2.3) by introducing �=e-0/=� O, , (2.2.4) the
Helmholtz free energy per unit length. By simple calculation, O O OO O 0 - Ot
O00t O(00/Ox) OxOt (2.2.5) O0/0 O0 -< q xx l - rl ' We want to ensure this will
hold for any choice of the function O(x,t), assuming (2.2.2) is used to calculate
q. Consider (x, t) = (Xo, to), any par- ticular point and time. It is easy to
construct temperature functions such that 0(Xo, to) = a, 0--0x (Xo, to) = b, (Xo,
to) = c, (2.2.6) 020 OxOt (x0, to) = d, where a, b, c and d are arbitrary constants,
with a > 0; simple polynomial functions will suffice. Now, use these values in
(2.2.5) evaluated at (x0,t0), to obtain an in- equality of the form A + Bc + Cd <
O, where A, B and C are functions of a and b only. Fix a, b, and c. If C > 0, we
could take d negative and large enough to violate this inequality and, if C < 0,
we could similarly violate it. A similar argument applies to B, so we must have
B=C=0, A<0, for all values of a and b, that is, for all values of 0 and O/Ox. This
gives 0 oo -- 0 = � = �(0), (2.2.7) () 2.2 Constitutive Theory for Rigid Bars 19
and 1= dO' (2.2.8) This leaves us with the inequality g 0,xx = qxx >0' (2.2.9)
Now, for 0 fixed, g clearly vanishes when O0/Ox = 0, and, being nonnega- tive,
it has a minimum there, so Og (0,0) = q(O,O) = 0. (2.2.10) 0 oo Physically, 0 is
constant in equilibrium and the heat flux then vanishes. Fourier's law (2.2.1) then
emerges in a natural way as a first approximation for O0/Ox suitably small.
Equation (2.2.9) reduces to the condition n k 0, which is always sumed in such
theory. With a more nonlinear theow, q must be chosen to satisfy (2.2.9). As w
mentioned in our considerations of equilibrium theory, it is com- monly assumed
that is a monotonically increing function of 0, so, by (2.2.8) d d2 - > 0, (2.2.11)
dO dO 2 and (2.2.8) can be inverted, to obtain 0 a function of . Also, from
(2.2.4) and (2.2.8), de de dr I _ o drl dO' involved earlier in (2.1.28), there
deduced from equilibrium theory. With this, the energy equation (2.1.8) becomes
Oe de O0 drl O0 - 0 ot dO Ot dO Ot (2.2.12) _ 0 0rl _ Oq + r. Ot Ox With, say,
Fourier's law and Newton's law of cooling, this gives de O0 0 (n(O) 00 ) dO Ot-
Ox xx + (03 - 0), (2.2.13) as the temperature equation, our discussion in Section
2.1 covering some possible boundary conditions. Linearizing this about 0 = 0a =
const., gives
20 2. Constitutive Theory of Heat Transfer for Bars and Plates a linear equation,
likely to be used in problem solving in cases where the change in temperature is
expected to be small compared to some constant value. The linearized equation
is of the form with 00 020 Co = no x 2 q- a(0, - 0), (2.2.14) de (Os), no = n(Os).
(2.2.15) Co = Equation (2.2.14), with a ---- 0, is known as the diffusion equation
or the heat equation. If (2.2.12) holds, we can calculate that dS dor q (; 00/02 dr-
dr vdx = + + q ox / j dx. (2.2.16) With this and the earlier discussion of heat
baths, it is fairly easy to con- vince yourself that the sumptions of reversibility
made in calorimetry cannot be quite correct according to this kind of
nonequilibrium theory. Generally, the nonequilibrium theory is considered to be
sounder. This and similar experiences with other nonequilibrium theories serve
to motivate the rather common opinion that reversible processes are not really
physically attainable. There is one ce in which a simple analysis gives an
estimate of the errors involved. Suppose we start with the bar in equilibrium,
with 0=0}=00 fort<0. (2.2.17) Suppose that we insulate the ends, so that q(L, t)
= q(O, t) = O, Dt. (2.2.18) Then, at t = 0, we change the heat bath temperature to
0 = (1 + a)00, (2.2.19) where a is a small number. Then, reonably, we can use
the linearized equation (2.2.14), with boundary conditions 00 a0=0 at x = 0, L,
and we can satisfy this by taking 0 independent of x. This would also be true if
we used nonlinear theory. However, with linear theory, we get a simple
equation, 00 a = - 0), = > 0. 2.3 Constitutive Theory for Thermoelastic Bars 21
Integrating this, and using the initial condition (2.2.17), we get 0s - 0 = (0s - 00)
exp(-/t) (2.2.20) = aOo exp(-3t), so 0 approaches 0s exponentially, as t - cx>.
The total heat supplied to the bar is 5Q=/ / rdtdx=aL (OB--O) dt (2.2.21) = L -
LCoaOo. With (2.2.16) the total change in entropy, according to nonlinear
theory, is AS = dt dx = aL dt. (2.2.22) As estimated by linear theory, we
calculate that Os - 0 a exp(-/t) 0 I + a[1 - exp(-/t)] a exp(-/t) with an error of the
order O(a2). In this approximation, we then obtain Laa 5Q AS -- - 3 0o' or the
approximate reversibility needed to correlate with equilibrium stud- ies such as
are involved in calorimetry. So, one needs/t to be large to get near
thermodynamic equilibrium and the heat bath temperature changes to be small,
as measured by a, to approximate results obtained from equi- librium theory.
This exemplifies a remark made in Chapter 1. The desired reversible processes
do not always exist. Often, we can approximate reversibility as closely as we
like, here by making a small enough. In the limit a = O, we do get a reversible
process, but it requires the bar to remain in equilibrium, at 0 = Os, and, for
calorimetry, we need the temperature to change slightly. 2.3 Constitutive Theory
for Thermoclastic Bars Here we generalize the previous theory, allowing the bar
to stretch. This involves picking a reference configuration of the bar. The usual
practice is to identify this as a rather stable equilibrium configuration, subject to
no
24 2. Constitutive Theory of Heat Transfer for Bars and Plates to obtain the
differential equation p - + o Oq -- f 9 + xx ( ag ) r xx ' Using (2.3.8), we can
cancel some terms giving O Oq (2.3.12) =o+r+o- , the idea being to use this as
an equation for O(x, t). Again, we use the Clausius Duhem inequality (2.1.9),
dSd x2 x2r dt -- dt 1 dX dX + , (2,3,13) merely reinterpreting S to be entropy per
unit reference area. As before, we then have 1(Oq) 0 r+ -q 02 . (2.3.14) The
usual idea is that, for a given bar, f and r may be specified in different ways,
depending on the systems with which it is to be in contact. So, we require a
theory enabling rather arbitrary specifications of these and we will not formulate
constitutive equations for them. In (2.3.12), f h already been eliminated and,
before, we can eliminate r between (2.3.12) and (2.3.14), to obtain the inequality
o o0 - oo + (q/O) . Also, before, it is convenient to introduce = - On, (2.a.lS) the
Helmholtz flee ener per unit reference volume, satisfying O Va + (q/O) O0 (2..1)
Then, we seek constitutive relations, relating , a, , and q to motion and
temperature, allowing for the fact that different bars can respond dif- ferently.
Here, we aim at theories which are, in a sense, very local. Consider any fixed
values of x, t, designated by x0 and t0. For small enough values of x - x0 and t -
t0, we have, approximately Oy y(x, t) y(xo, to) + (x0, to)(x - xo) + (xo, to)(t -
to), (2..x7) o0 O(x, t) O(xo, to) + (x0, t0)(- x0) + O(xo, to)(t- to). 2.3
Constitutive Theory for Thermoelastic Bars 25 Then, roughly, the assumption is
that the behavior of a material point is influenced by the general motion and
temperature. This influence is small except when it occurs very close to the point
in position and time. Using (2.3.17) to estimate this, we are led to the assumption
that b, a, , and q are functions of the following quantities: Oy O0 y, A--0' ' 0, xx' '
(2.3.18) Two arguments are used to restrict these functions. One is the notion
that these should be objective functions, as it is sometimes put, meaning that
their values are unaffected by superposing rigid motions. Suppose, for example,
that we have two motions y(x, t) and (x, t), with (x, t) = y(x, t) + a + with a and b
constant. Here, it would be acceptable to replace a + bt by an arbitrary smooth
function of t, but one gets the same restrictions, either way. Suppose also that the
temperature functions are the same, (x, t) = 0(, t). It is true that a given particle
takes on differcnt positions and the velocity differs by a constant; but,
physically, we do not expect this to affect heat fluxes, entropies, and so on.
Thus, for cxamplc, we should have Oy oo y-a+bt, xx' -b, O, xx' ( Oy oo) = y, , ,
0, , O , for any given motion y(x, t) and any choice of constants a and b. It then
follows that this will be true only if b does not depend on y or . By this kind of
argument, we reduce (2.3.18) to Oy O, 00, . (2.3.19) 0-' and q depend on these
variables, we put this into Assuming q, /, a, (2.3.16) to get 1o (o,)o,.. (2.3.20) O
O (q/O)-. + O(00/Ox) Ox - As before, we want to restrict the constitutive
functions so that the in- equality is satisfied for all motions and temperature
distributions. We use
26 2. Constitutive Theory of Heat Transfer for Bars and Plates the fact that, with
the variables (2.3.19) fixed, at x -- x0, t = to, we can still regard as independent
variables higher derivatives such as . This gives the restrictions (2.3.21) 04, 04,
O(Oy/Ox) with the remaining restriction that (q/O) - + vo. Since this expression
vanishes when = OO/Ox = 0, it then takes on its minimum value. Applying the
first derivative test for a minimum to this function then gives = oo/Ox = 0 q = 0,
v = -oz/oo. Usually,. workers sume that q is independent of and that V is
indepen- dent of and OO/Ox, and we will follow suit. With the indicated
sumption we have, always, V = -0/0, q , , 0. (2.3.24) Often this is specialized
more, with the sumption that q=k , O 0' k0, (2.3.25) which seems to be adequate
practically. Underlying (2.3.24) is the classical view that entropy is defined
using processes which are, at let approx- imately, reversible. It would not hurt
for you to think hard about this yourself, but it does at least seem dicult to
reconcile this with the idea that V can depend on and/or OO/Ox. The effect of
this is to sume that, at each particle, the entropy is given by the constitutive
equation applying to equilibria. For various other theories which are, in a similar
sense, local, therm dynamicists often use this sumption with some success. It is
called the pnnciple W local states or the hypothesis W local theodynamic
equilibum and other similar names. Often workers interested in viscoeltic effects
em- plw theories which are not local, in the temporal sense. For example, the
�If you wish to explore this more deeply, you might find it helpful first to study
Chapter 10. 2.3 Constitutive Theory for Thermoelastic Bars 27 stress at any
particular time depends on values of the stretch experienced at all previous
times. Here, such assumptions are commonly rejected. For example, one expert,
Rivlin [12] offers physical arguments to help explain why he considers them to
be unsound for use in connection with such the- ories. Various other experts
have made clear that they came to a similar conclusion. This will be discussed in
more detail in Chapter 10. It is not reasonable to infer from this that the reasons
that convinced one would be the samc as those that convince another. Obviously,
one needs some alter- native, and, concerning this, there are some differences of
opinion. There are still different ideas about entropy, stemming from statistical
molecular theories, for example, which do not require the association with
reversible processes. Such theory also suggests a need for generalizing the
Clausius Duhem incquality. It would take a very lengthy discussion to elaborate
these remarks so we will not pursue it. With the specialization indicated, we
have what is generally regarded as thermoelasticity theory or, more properly, the
one-dimensional version of such theory. Such theory has been used,
successfully, to describe numerous phenomena. With a given by (2.3.21), the
equations of motion (2.3.8) become /) P=xx +f' (2.3.26) it being understood that
some appropriate prescription for f is to be given. In most situations, it is
reasonable to take f = 0. When exceptions are en- countered, they are likely to
involve forces exerted by gravitational or elec- tromagnetic fields and, in the
latter case, it is probable that more general kinds of constitutive equations are
needed. To obtain the temperature equation, we use (2.3.12) and note that with
our constitutive equations, oV=-o from which (2.3.27) Superficially, this is the
same as the equation (2.2.12) derived for rigid bars. However, here q depends
not only on 0, but also on A, so (2.3.26) and (2.3.27) are a coupled set of
equations for motion and temperature. Again, r needs to be specified and the
possibilities for this are well-covered by our discussion of rigid bars.
30 2. Constitutive Theory of Heat Transfer for Bars and Plates the line now
being normal to the plate in its reference configuration. Motion involves a
displacement u, in a perpendicular direction, described by an equation of the
form u = u(x, t), (2.4.2) the analogue of the previous y. With it is associated the
velocity, Ou /t = , (2.4.3) and the analogue of the stretch A, the shear strain Ou '
= xx' (2.4.4) with - -- 0 in the reference configuration. There is a difference in
that, physically, A should be positive, while - can be positive or negative. As-
sociated with the shearing motion is a shear stress r analogous to a. The
equations of motion are of the form 0r p5 = xx + f' (2.4.5) f acting in the
direction of the displacement u or its opposite. Then, as analogues of (2.3.21),
(2.3.24), and (2.3.28), we have relations of the form or 0 = 0(v, 0), (2.4.6) oo r/-
00' (2.4.7) ov For a temperature equation, we have the copy of (2.3.27), Oq 0//=
xx + r. (2.4.8) One difference is that here a notion of material symmetry is of
some importance. For plates made of isotropic and some anisotropic materials, it
is reasonable to assume that q5 is an even function of % qS(h,, 0) = qS(-h,, 0).
(2.4.9) 2.5 Thermodynamic Experiments 31 Reversing the sign of 'amounts to
reversing the shear displacement. For, say, isotropic plates, one expects that this
will be associated with forces of the same magnitude but opposite directions (r -
-r and f - -f), keeping O(x, t) the same. So, (2.4.9) summarizes common
experience of this kind. One can encounter cases where it is unreasonable to
assume (2.4.9), in dealing with some kinds of crystal plates, for example. Later,
(2.4.9) will play an important role in discussions of Martensitic transformations,
phenomena which are rather common in crystalline solids. As is rather obvious,
various ideas and results used in analysing bars can be borrowed to analyze
plates and vice versa, and we will not belabor the obvious. 2.5 Thermodynamic
Experiments Clearly, it is important to know the form of the function q5 for
particular bars or plates and there is a rather common strategy for using experi-
ments to try to accomplish this. The same kinds of ideas are used in deal- ing
with three-dimensional thermoelasticity theory. Later, we will consider some
two- and three-dimensional theories. To be definite, we will discuss the bars. We
would like to induce the bar to be in various configurations in which A and 0
take on a variety of values, independent of position. Physically, we would like to
know that each of the pairs of values can be connected to the reference values (A
= 1, 0 = 0R) by a process which is, at least to a good approximation, reversible.
Theoretically, entropy changes are involved in a rather important way and we
need some assurance that it is physically well-defined. Commonly, workers try
three types of experiments: 1. measurements of thermal expansion; 2.
measurements of specific heat Co(O) in unstressed specimens at vari- ous
temperatures; 3. static mechanical experiments performed at constant
temperature, at a variety of temperatures. If all goes smoothly, this combination
can, in principle, suffice and for a first look we will assume that it does. Thermal
expansion refers to the deformation caused by a change of tem- perature in an
unloaded body. Theoretically, we would derive it by solving for A, oo er = (A,0)
= 0, (2.5.1) to get A = X(0). (2.5.2)
38 3. Equilibrium Theory of Bars For a static analysis we are not concerned with
motions but deformations y = y(x). However, we can allow any which could
occur instantaneously in a motion. A similar remark applies to temperature
functions O(x) or /(x). The condition that the upper end is clamped limits the
possibilities. We represent this as the condition that y(0): 0. (3.1.1) This in turn
means that no work will be done on this end. From a dynam- ical point of view,
power is supplied to the other end by the fixed force, expressible in the form P =
k(�, t): -x/t (3.1.2) where X = -ky(L, t) + const., (3.1.3) k being a constant
measuring the size of the force. It is convenient to take this as the weight divided
by the reference cross-sectional area of the bar since we divided the bar energy
by this area. From (1.1.4), we have here our first example of a conservative
loading device. For static analysis, the dependence on t is irrelevant. In principle,
the bar's own weight will cause it to stretch a little; very little for the small bars
used in laboratory experiments. So, as is customary, we neglect such effects,
assuming that the body force f -- 0. Of course, it is understood that forces are
applied to the ends of the bar and not to its sides. Then, from the discussion in
Chapter 1, we expect the system to approach equilibrium, our situation being
covered by statement IV, with E here interpreted as the energy of the bar. The
relevant thermodynamic potential is then given by (1.3.5) or E x =E-OsS+x
(3.1.4) L = fo ( - OBI) dx - ky(L) and we are concerned with its variations. We
interpret this in much the same way as we did in the consideration of rigid bars.
For bar theory, we should have some constitutive equation, it being con- venient
to use one of the form Oy = (, 1), - Ox' (3.1.5) This has the features noted in
(2.3.28). First, we want to know what condi- tions y = (x) and /= (x) must satisfy
in order to qualify as equilibrium configurations. All deformations must satisfy
(3.1.1). So, we consider func- tions of the form y(x) = + (3.1.6) v(x) = + 3.1
Equilibrium of Bars Subjected to Dead Loads 39 where/ is a small parameter and
the pair (Sy, &/) are any smooth functions of x such that 5y(0) = 0. (3.1.7) This
describes variations in the system. As before, we calculate the variation in E x by
regarding 5y and &/to be fixed, so E x reduces to a function of/ only, and taking
5E x = Ex(0 ). (3.1.8) Then, statement IV is interpreted as meaning that we
should have i. tSE x _> O, (3.1.9) for any admissible choice of 5y and 5/.
Assuming that all functions involved are smooth, we can differentiate under the
integral sign to get ( )] 6E x = - -- + - OB 51 dx - k6y(L), (3.1.10) Ox with
overbars denoting evaluation at the putative equilibrium state. From the general
theory, the stress P and temperature in this state are given by c _ c (3.1.11) We
then perform an integration by parts to replace O(Sy)Ox by 5y dx = Sy - xx 5y
dx, a standard procedure in such variational problems. So, using (3.1.7), we have
5E x = ( - O)&l - xx Y] dx 4- ( [= - k) 5y(L) (3.1.12) and interpret the criterion
for equilibrium as the condition/SE x _> 0. Note that, if/SE x were positive for
some choice of (&/, 5y), it would be negative for 5- = -&/, 5- = -Sy, another
allowable choice, so we must in fact have 5E x = 0 for all variations:
equivalently, we could take/ positive or negative. By arguments similar to those
used for rigid bars, we deduce that the equilibrium conditions are that 0 0 ==>
const., (3.1.13) Ox
74 5. Balloon Problems Then, it gets harder again until the balloon decides to
burst. It is hardly reasonable to use the same analysis for long tubular balloons,
although one observes somewhat similar phenomena, along with aneurisms
which develop and grow in length as air is added. To simplify analyses, we
assume that the balloons stay spherical. Insta- bilities could occur, inducing the
balloon to take on some less symmetrical shape. Later, we will discuss some
evidence suggesting that the possibility is real. However, it is a more difficult
problem to analyze this, and, for the experimentalist, it is hard to get balloons
that are very good spheres of uniform thickness, and so on. It is a matter of
experience that, near condi- tions at which instabilities occur, small differences
of this kind can have a significant effect. Such quirks are discussed by
Thompson and Hunt [25]. To get an idea of the number and character of
equilibria, it helps to consider another picture. First consider the graph of
OF/OV, regarded as a function of V. This differs from that given in Fig. 5.2 only
by a simple change of scale. For simplicity, assume that we use the ideal gas
model, described by (5.1.11) with Vc -- V. With V as abscissa and p as ordinate
add in the p-V curves, in this case hyperbolae, for various values of M. From
(5.2.4), the possible equilibria are given by the points where one of these
intersect the graph for the balloon. From such a sketch, one can see that there is
at least a theoretical possibility that by adjusting M, one might get one, two or
three equilibria? Consider the picture in Fig. 5.3 when there are three, giving V <
V2 < V3 as equilibrium volumes. Now, as we did before in considering the equal
area rule, we can picture energy differences as area differences. Obviously - pdV
- - pdV (5.2.6) with the latter integrals referring to the gas. However, with (5.1.9)
we can replace these by differences in Fc, for example -/v2pdV---F6()-F6(-) �
(5.2.7) 2By adjusting the shape of the rubber curve, as indicated in Fig. 5.3, you
can get a rather similar sketch where the two curves never intersect more than
once. This is a way of constructing cases where (5.2.5) always holds. According
to Alexander [26], this type of behavior is encountered in some neoprene
balloons used for high-altitude measurements. 5.2 Equilibrium of Balloons
Containing a Fixed Mass of Gas orp 75 v. v, v v v FIGURE 5.3. Sketch of the
balloon and gas response curves, for a value of M giving three points of
intersection. Here A and A2 denote the areas of the cross-hatched regions. The
dashed curve indicates another possible graph for the balloon giving only one
point of intersection with the graphs for the gas. Then, with (5.2.2) we have
E(V2, M) - E(Vx, M) -- Az > 0 (5.2.8) E(V2, M) - E(V3, M) = A2 > O. Thus,
the equilibrium at V = V2 is less stable than those at Vz or V3. It is in fact
unstable, as can be seen by doing similar area comparisons, for V near V2. The
stability enhancement indicated by (5.2.5) is associated with the possibility of V
taking values where oq2F/oqV 2 < 0, as it can. Again, area comparisons indicate
that, as long as we have the three equilibria, V and V3 are both at least relative
minima. From (5.2.8) one can read off which is stable: A1 _ A2 V1 stable,
(5.2.9) Az _< A2 = V3 stable. So, we do have an "equal area" rule somewhat
similar to those discussed earlier, although the areas are not defined by
intersecting the basic graph with a straight line (p = const.). With a single
spherical balloon, we do not have a physically reasonable analogue of the phase
mixtures encountered in bars and plates. One begins to obtain something similar
if one considers several balloons, interconnected so gas is free to move from one
to another. Here, since pV is proportional to M for an ideal gas, increasing M
moves the p-V curve outward, in an obvious sense. If M is small enough, it will
intersect the graph just once, to the left. As M increases, it will attain a value
where the p-V curve touches the graph: here V2 and V3 just appear, with V2 =
Va. As M increases further, these split, V2 becoming unstable, Va metastable.
Further increasing M brings us to the equal area configu-
6 Biaxial Stretch of Rubber Sheets 6.1 The Idealized Problem We now consider
biaxial stretch experiments performed on rubber, one of the isothermal
mechanical experiments done to obtain information about the relevant strain
energy functions. Again, we choose as a reference a stable unloaded
configuration, say at room temperature. For simplicity, we will only consider
experiments at this temperature. In the reference configuration, the sheet
occupies a region described by -a _ X 1 __ a, --a <_ x2 <_ a, -b _< xa _< b,
(6.1.1) where (X l, X2, X3) are coordinates in a suitable selected rectangular
Carte- sian coordinate system. Generally, b/a is quite small; thin sheets are com-
monly used. The aim is to produce simple kinds of deformations by apply- ing
suitable loads to the edges of the sheet. Briefly, the material point at (Xl, X2,
X3) goes to (y, y, Y3) with yl = ,1Xl, y = A2x2, Y3 = '3X3, (6.1.2) where the
stretches Ai are positive constants. That is, we will only consider deformations
of this kind. Here, the assumption is that the strain energy function W is a
function of Ai, satisfying W(l, 2, 3) = W(2, 1, 3) = W(l, 3, 2) (6.1.3)
84 6. Biaxial Stretch of Rubber Sheets 6.2 The Treloar Instability 85 You may
explore for yourself the other implications. becomes E = V - r5 - r,7, where =
+T2, r=T1-T2, and the equilibrium equations now become Also, (6.1.10) now
(6.2.4) (6.2.5) ov ov r = 0(i' r = 0,7 (6.2.6) With V an even function of ,7 and r
related to it as the derivative with respect to '7, r and ,7 become analogous to the
shear stress and shear strain considered in the theory of Martensitic
transformations. That another vari- able (i is involved does complicate matters
somewhat. However, its varia- tions may be considered to be similar to the
variations in 0 considered before. In these terms, Treloar's data indicate that, for
some values of (i, r = 0 but ,7 0, giving us an analogue of Martensite. It is not
likely that this is true for all & For example, indications are that there is only one
unloaded state, so r = r = 0 = (i = 1 and ,7 = 0, giving us an analogue of
Austenitc. Indications are that something more or less like a Martensitic
transformation should be encountered in some loading programmes. One might
expect to see something similar to twinning but experimentMists seem not to
have reported this. Theoretically, one can understand this. If one tries to
construct deformations involving jump discontinuities in '7, even allowing them
in (i, one finds that it is impossible to have all three components of displacement
continuous. Thus, in this respect, the analogy with the theory of Martensitic
transformations breaks down. Conversations with some rubber experts indicated
they were unaware of the possibility of such an instability, until Kearsley
revived the issue and his work is not yet widely known. Occasionally, something
small can produce a surprising error in an ex- periment, so it is important to try
to make an assessment of this situation theoretically. One approach is to take a
form of W which can be analyzed and fits some relevant data, at least roughly.
For rough calculations, work- ers often use the Mooney-Rivlin form, W
C(A12+A2+A32 3)+C2(12 1 1 ) : - + , with Ct and C2 constants, determined by
curve fitting. What is important for our purposes is their ratio, so we will use the
simpler form I 1 1 W = K(A 2 + A22 + A32 - 3) + A + A + A - 3. (6.2.7) This
gives V=K [2((i 2+,7 2 )+ ((i2 _ '772)2 1 1 + ((i +,7) + ((i - Then a calculation
gives + ((i2 _ '72)2 OV 4'7 0'7 ( (i2 _ '72)3 (6.2.8) = {K[((i2-,72)a+l]+,72+3(i2-
((i2-,72)4}. (6.2.9) One of the conditions for metastability is that 02V/0,72 _> O.
Let ns explore this at '7 = 0, using 02V' = lim _1 o___V_v 0'72 ] =o -o ,7 0,7 =
if6 [( K((i6 + 1) + 3(i 2 - (i s] (6.2.1o) _ 4((i 6 + 1)[K - f(&l, where (i2((i6 _ 3)
(6.2.11) if(i)- (i6+1 ' is rather easy to visualize, from its graph. For a positive
value of K, such as that indicated in Fig. 6.2, there is just one value of (i = for
which f = K, and we have K > f((i) for (i < , K < f((i) for (i > , so we have
instability at ,7 = 0, when (i > , and when (i < the one stability inequality is
satisfied. Curve fitting produces various values of K depending on the particular
rubber and the range of deformation which one selects to fit best. Kearsley
mentions that likely values range from 4 to 8, although one can find estimates
outside this range. Roughly, this puts somewhere between 2 and 3, not
unreasonably large for rubber. Also, this means that for any fixed (i < and '7
close enough to zero, V is a convex function of '7 and this is not true for (i > .
Thus, for fixed (i < , r is a monotonically increasing function of '7, suggesting
that we have an analogy with the second-order Martensitic transformations. To
consider other possible solutions of r = 0, we note that they will occur when
K[((i 2 - ,72) a + 1] + '7 2 + 3(i 2 - ((i2 _ '72)4 = 0.
102 7. Moving Discontinuities one would need to say more about the mechanism
for applying the force and we will not pursue this. Otherwise, the model is
reasonably consistent with experience. Let us try to obtain a better understanding
of what it predicts. If, as was assumed, W(A) is a convex function of A, a() is a
monotonically increasing function of , with a(1) = 0. Thus the values of
satisfying (7.3.5) will, for positive f, increase with f and be greater than one.
Consider, for any fixed value of g() = b[W(A) - ( - cos )o-1 + a, (7.3.9) we have
and 9(1) = aGa > O, (7.3.10) g'(A) = .b[W'(A) - - (A - cos (7.3.11) = -ab(A - cos
0)cr' < 0. For A > 1 and (A- 1) sufficiently small, g(A) will be positive by
continuity and this will correspond to having f small enough. Then, (7.3.7b)
implies that the whole tape should remain in contact with the other body. As f
increases, (7.3.5) implies that g(A) decreases. Mathematically, it could approach
a positive constant as A becomes large. Physically, something must happen if the
force gets sufficiently large but there is some possibility that this may occur, by
breaking out chunks of the other body, for example. Assuming g does continue
to decrease, it will vanish at some critical value fc of the force. Here, in
principle, any part of the tape could peel off with the rest staying in contact. For
f > fc we will have g < 0, the somewhat shaky (7.3.73) seeming to imply that the
whole tape should come loose. With the inevitable experimental errors, it is
unlikely that we will attain the value of fc exactly. Granted this, one should see
the tape either staying in complete contact, or coming completely free,
depending on the size of the force, in this kind of experiment. This is consistent
with experience on relatively long tapes, for which L/a is relatively large.
Generally, applying a force in the manner indicated either leaves the whole tape
in contact with the other material, or the whole tape comes off. If a force induces
it to come off, a larger force simply makes it peel off more quickly, assuming
that there are no other complications such as having the tape break. Also, from
our considerations, the critical force fc occurs when (7.3.8) holds and this clearly
depends on the angle 0, which should be in the range 0 physically. In this range
and with cr 0, as it must be, it is easy to see that for any fixed value of A ) 1, g
increases with T. Combine this with (7.3.11) and consider the critical values of
A, say Ac(0), obtained by solving g(, V') = 0, (7.a.1.) and you can conclude that
A must decrease as o increases. From this, it follows that fc is a monotonically
decreasing function of o. At least qualitatively, this is also in accord with
experience. 7.4 Another Peeling Problem 103 Another picture can be useful.
Rewrite (7.3.8) as (/k c -- COS (fi)cr(.c) -- W(/kc) ---- Ga/b. (7.a.la) Then,
consider the graph of or(A) in the A-or plane. The first term can be interpreted as
the area of a rectangle with height cr(Ac), base (Ac - cos o), the length of a line
segment on the A axis, running from A = cos o < 1 to A = A 1. From this, we
subtract W(Ac) which we know is the area under the stress stretch curve
between A = 1 and A = Ac. Thus, the left side of (7.3.13) represents the area of
the rectangle which remains after removing the latter part. When this area
matches the number on the right, the equation is satisfied. From our previous
considerations, there cannot be two values of Ac satisfying it if W(A) is convex.
If it is not, then one can get into complications such as are discussed in Chapter
3. 7.4 Another Peeling Problem Here, we consider a procedure permitting us to
peel off only part of the tape, also indicating how we might apply the tape so
that, after it is placed, it will be unstressed. Again, we gloss over some physical
difficulties occurring when only a small part near one end is in contact. Here, the
end x = L is to be moved perpendicular to the rigid body to a height h, then held
at this fixed position. Otherwise, the assumptions are similar to those made
before. Figure 7.3 indicates the configuration to be considered. Here, the angle o
is not given. As before, W(A) is assumed to be a convex function for simplicity
and we again ignore complications associated with damage to the materials. It is
then rather clear from our past experience that, in the part pulled off, the stretch
A will be constant, so we will assume this although it is something that could be
proven. Here, with the end held fixed, no work will be done on it so, in place of
(7.3.4), we now have L F=,o f w(A)cx +G.(�-xo)a o = a[bW() + (7.4.1) where
and note, from Fig. 7.3, that I = L - x0, (7.4.2) d'2 = 12 + h 2 (7.4.3) ____ ,2/2,
104 7. Moving Discontinuities / d � h FIGURE 7.3. Raising the end of the tape
vertically, to a height h, partially peels off the tape. and tan qo = hi1. (7.4.4)
Thus, two of the three variables (A, l, q) can be expressed in terms of the third.
We could take any one as the basic independent variable, it being a matter of
judgment which is best. First, let us try using = l/h 0 _< _< L/h. (7.4.5) Then,
(7.4.3) gives A = X/f+/z -2 > 1, (7.4.6) and (7.4.1) reduces to the form where
F/ah = U(.) + (7.4.7) (7.4.s) If/z is very small, A is very large, approaching
infinity as p -+ 0. Physically, it is then unlikely that we will have an end point
minimum at p -- 0. Also, having p at the other limit really means that all the tape
h peeled off. If we continue to hold one end fixed the remainder will be subject
to no force, so no work will be done on the tape. It should then come to
equilibrium, unstressed, giving F = aLC (7.4.9) a reonable estimate of the energy
it will then have. The remaining possibility is to have equilibrium with not at
either limit. Then, math- ematically, analysis of (7.4.7) is essentially the same
that of analyzing bars under dead loads discussed in Section 3.1, or of the
balloons under fixed pressure discussed in Section 5.1. The equilibrium equation
is '() = -C, (7.4.0) 7.4 Another Peeling Problem 105 the second derivative test for
stability giving U"(p) k 0. (7.4.11) Conditions for stable equilibria can be
pictured, as before, in terms of the graph of U(/) or of U'(I). Some points are
worth noting. Suppose (7.4.10) is satisfied by some value of/ -- . This is
acceptable only if it satisfies < L/h, which will be true if we make h small
enough, but not if h is too large. For the range of h for which it is acceptable,
(7.4.4) will give a value of qo = which is independent of h. Ignoring the question
of stability, the implication is that if we increase h a little, more of the tape will
come loose, enough to get back to its original value with no change in the value
of Clearly, it can so adjust only as long as there remains enough tape to be
peeled off, that is for h <_ Ltan. (7.4.12) For such configurations to be more
stable than they would be if the tape pulled off completely, using (7.4.7) and
(7.4.9) gives the condition h[U() + G] < LG. (7.4.13) Now consider (7.4.8). With
(7.4.6), straightforward calculations give b + =b [W(A)-(-a 1) -a(A)] (7.4.14) :
Note first that, from Fig. 7.3, L - x0 I 1 ..... . (7.4.15) cos - d d With this, we see
that (7.4.14) agrees with (7.3.8), which is not surprising. Also (7.4.16) _ + d
Now, kom (7.4.6), dA/d < 0 and A > 1. om our smption that is convex with (1) =
O, it follows that (A) > 0 for A > 1 and (A) k O, kom which U"() > O. (7.4.t7)
106 7. Moving Discontinuities There then can be no more than one equilibrium
value of the kind dis- cussed above. Also there can be at most one value of the
angle , in the range 0 <_ < r. Assume there is one. Then, for h satisfying (7.4.11),
we have equilibrium configurations satisfying all the requirements for stability
except perhaps (7.4.12). With a fixed number, this will also hold if h is
sufficiently small. For the largest possible value of h, given by the equality in
(7.4.11), it is easy to see that (7.4.12) fails to hold. Thus the two energies
become equal at some value hc, given by hc = ' (7.4.1s) So for h < he, the more
stable configuration has the tape partially adhering, enough to conform to the
angle . For h > he, the more stable configuration changes, to have the tape
completely removed. From the analysis, the pos- sibility of the tape adhering
when h > hc giving a metastable configuration does not seem unreasonable in
cases where we are increasing h, to peel off more of the tape. In laying a tape
down, we commonly hold the part near an end in place to get the process started.
Clearly, our analysis does not cover this. Essentially, the theory is designed to
apply to situations such that one can lay down and peel off the tape without
damaging the tape or the other material. To determine W(A) for A 1, most would
try a simple tension experi- ment. Either of the two peeling experiments could be
used to try to estimate Ga. Here, one could measure the stretch in the stretched
part. Similarly, hc might be measured. Think a little more about the measurables
in the two experiments and you have a design for an experimental programme to
provide a test of the theory in order to see if its predictions agree with the
experimental findings. This is a sample of ideas used in tackling problems
relating to the ad- hesion of one material to another. Works dealing with this
general topic include the books by Cheng [36] and Wu [37]. 7.5 Exercises For
Exercises 7.1-7.3, adapt the theory of fast shock waves in bars to the theory of
shearing of plates, using the constitutive equation �_ a(1-bO) 2 cIOln(- )_O+d
1 2 ' where a, b, c, and d are positive constants. For various metals, empirical
estimates of b give it the value 1/(202u), where OM is the melting temper- ature,
so the theory does not apply if 2bO _> 1. If some prediction violates this, note it.
For the exercises indicated, assume that, on one side of a shock, 7 = O, 0 = d,
and on the other, 7 = � > O. 7.5 Exercises 107 7.1. Derive the Rayleigh
equation. 7.2. Derive the Rankine-Hugoniot equation. 7.3. What can you say
about the admissibility of such a wave? 7,4. Consider the kind of peeling
problem discussed in Section 7.2 for a linear elastic tape with the strain energy
function W = E(A - 1)2/2, E being Young's modulus, a positive constant. Let
f(b) denote the lower bound of the force required to peel off the tape, when it is
applied at the angle b. Derive formulae for the ratios f()/f() and f(3_)/f(). 7.5.
Derive an equation relating isothermal and isentropic acoustic wave speeds for
bars, in terms of things either measured or easily calculated from things
measured in the common thermodynamic experiments.
8 Mixture Theory 8.1 General Remarks From a macroscopic point of view, the
composition of matter can change in various ways. It is a familiar fact that
changes in the humidity occur in the air we breathe, causing wood to shrink or
swell. This makes it reasonable to think that the wood absorbs different amounts
of the vapour depending on its environment, producing an effect somewhat like
thermal expansion. Materials like glass are clearly more reluctant to absorb the
water, making it clear that the water vapour is not uniformly distributed in a
room con- taining various kinds of materials, although it may be uniformly
distributed within a particular substance. Again, it is a matter of experience that
different amounts of a solid can be dissolved in a given amount of a suitable
fluid and that, if we change the temperature, say, some of the dissolved solid can
come out of solution. Sometimes we use this to grow larger crystals from small
ones. Not so different are the procedures used by a metallurgist to make alloys;
that is, melt and mix together different kinds of solids, then cool the mix to
solidify it. With the same ingredients mixed in different proportions he can
obtain different alloys. For example, he might produce a or /3 brass by using
different proportions of copper and zinc. In many situations like this, it is
reasonable to think that the ingredi- ents retain their identity so the total mass of
one remains fixed. We are then concerned with mixtures which are, in a sense,
nonreacting. Chemical reactions can change the mass of particular ingredients
with, for example,
110 8. Mixture Theory hydrogen and oxygen gases reacting to produce some
water. Stoichiome- try provides rules, restricting how the different masses can
change in such reactions. Gibbs' general ideas of thermodynamic equilibrium
and stability can be applied to such phenomena. Indeed, Gibbs was a pioneer in
developing such theory. Here, we will discuss the most elementary format for
nonreacting mixtures, in situations where effects of shear stresses in the solids
can be ignored. To deal with such questions one needs to think of equations
describing parts of a system and this involves some judgment. In the first
example mentioned, most would think of the air and water vapour as one mate-
rial, described by one constitutive equation. A different equation would be used
for the wood and its water vapour. Generally, one idea is to try to ar- range that
each constitutive equation can reasonably be considered smooth enough to let us
use common tools of analysis in each part. As we have seen before, smooth
constitutive equations can produce equilibria with dis- continuities and we tend
to judge things on the basis of observations of equilibria. Given this, it is
inevitable that we exercise some judgment. In the simplest situations, one can
reasonably decide how to further subdivide so equilibria are also smooth in each
part. 8.2 Elementary Theory We consider a hypothetical example to introduce
some of the ideas. Pic- ture a mixture filling some volume V. Consider it to be
composed of ho- mogeneous parts. That is, within a part, relevant state variables,
such as temperature, etc., can be considered to be independent of position. So the
presumption is that we know how to subdivide so that the equilibria are smooth
in each part. The parts will occupy volumes with =v. (8.2.1) i=1 An identifiable
ingredient, such as our water vapour, may be present in all of these parts and we
allow that this can happen, initially, and that the amount in one part can be
considered to change. Suppose that we have n ingredients, with masses ME, a =
1,..., n. Then, if M (i) is the mass of the ath ingredient in the volume V/, we must
have M (i) = ME, a = 1,..., n, (8.2.2) i=1 8.2 Elementary Theory 111 it being
possible that some of the M (i) vanish. This is a static theory, so we are not
concerned with motion. With the ith part we associate a total energy Ei and
entropy $i and a constitutive equation. That for i = i is assumed to be a smooth
function of the form Z 1 = Ei(Vi,Sl,M?),...,Mn(1)), (8.2.3) with similar
assumptions for E2, etc. It is possible that one constitutive equation applies to
two more parts. The assumptions are fairly reasonable if we have solid parts
separated by fluid parts, so the solids can adjust their shape fairly freely to avoid
shear stresses. The idea is that the partial volumes and masses can vary, subject
to the requirement that (8.2.1) and (8.2.2) be satisfied; we do not want
overlapping volumes, for example. We set OEi (8.2.4) Pi -- 0V/' interpreted as
the pressure in the ith part, (8.2.5) Oi - OSi ' the temperature in this part and OEi
ti, OM(i) , (8.2.6) these being called chemical potentials. Now look at the
differential of Ex. One term, -p dV can reasonably be interpreted as the work
done by the pressure as the volume V1 undergoes an incremental change. An-
other, OdSx, similarly can be identified with heat supplied reversibly. If we
stopped here, things would make reasonable sense in terms of the first law.
However, there is a problem in dealing with the changes in E pro- duced by
changing the masses in V which does not really fit the first law. One could try to
generalize the first law to cover this. What is more often done, sometimes,
tacitly, is to consider that any thermodynamic system deals with a fixed set of
matter. From this viewpoint the energy E is then an energy associated with part
of a thermodynamic system, which is not itself a thermodynamic system. The
usual idea that energies and entropies are additive then gives for the energy E
and entropy $ of the system, m E = y]Ei, i=1 m S = y]Si. i=1 (8.2.7)
114 8. Mixture Theory where the dV/ occurring here are arbitrary, so we must
have Pl P2 ..... Pm, (8.2.16) the equality of pressures which we might also infer,
mechanically, as a balance of forces. Similar consideration of variations of
masses gives, with (8.2.13), the equality of chemical potentials given by Pa =
P2a ..... Pm, a = 1,...,n. (8.2.17) So, (8.2.16) and (8.2.17) give the equations of
equilibrium. Of course, we also must satisfy (8.2.1) and (8.2.2). With Ms and V
here considered as given, the number of equations is then the same as the
number of unknowns (V/ and M(i)). While we shall not discuss in detail the
equilibrium conditions which apply when some of the volumes and/or masses
vanish, there is another assumption concerning constitutivc equations which
deserves to be men- tioned. Briefly, it is that, if we multiply all masses and
volumes by a com- mon factor, we can get the new energy by multiplying the
old value by the same factor. Roughly, it is the proportions which matter, not the
absolute amounts. For example, for the function F, we assume that for any
positive number k, and any possible values of the arguments, we have a (kV, 0,
kM?) -- 1 (V, 0, M?). (S.2.18) as a restriction on the form of this function. If we
differentiate this relation with respect to k, then set k = 1, we obtain the identity
n -pig1 q- Z I/laMa(I) -- Fl' (8.2.19) There are various ways of taking care of this
restriction. For example, taking k = l/V1 in (8.2.18) and setting Fl(X,O,p(a 1)) :
(1(0, p(al)), p(a 1) : MI)Iv1, we have Fx = Vq (O, p?), (8.2.20) are interpretable
as mass densities, qb as the Helmholtz free where p(l) energy per unit volume.
As is easy to check, if we take any function (/)1 of the indicated arguments and
use (8.2.20) to define F, it will satisfy (8.2.18) and (8.2.19). This opens the door
for dealing with problems in which the mass densities p(l) might vary with
position, as they will in a gravitational field. This can be important for
estimating water content in soil near a deep lake, for example. Then, the total
Helmholtz free energy F in a region occupied by this component could be taken
as F ---/�1(0, p(2 )) aV, (8.2.21) 8.2 Elementary Theory 115 giving (8.2.20)
when the integrand is constant and the volume of the region is V. Clearly, much
the same considerations apply to the other Fi and to the El, with the
understanding that entropies are to be replaced by entropies per unit volume.
With (8.2.20), one can verify that (8.2.19) reduces to -p + /aP ) = (/)1' (8.2.22) A
slightly different formulation can be obtained by introducing m (1): Ma (1),
(8.2.23) the total mass in V and the concentrations (mass fractions) c( 1) =
M()/m 0), (8.2.24) numbers lying between zero and one, satisfying n c( 1) = 1.
(8.2.25) ot=l Then, one can introduce the specific volume v = V1/m O) 1 = --,
(8.2.26 Pl where/91 is the total mass density in Vx. Then, taking k = l/m0), we
see that we can also write F in the form F 1 : TY/(1)()l, 0, c 1)) : Vlpl,()l, 0, cl)),
(8.2.27) and we can, if we like, use (8.2.25) to eliminate one of the c(a 1).
Originally, Gibbs introduced this kind of theory as a theory of fluid mixtures but
later workers, like those interested in alloys, have applied it to cases where some
of the parts are solid. Confining solids to a fixed region can easily induce shear
stresses, so it is a little more natural to think of V as the volume of a region
partly occupied by solids which are surrounded by a fluid at a fixed pressure p,
where the shape and volume V of the region can change, to help avoid such
complications. Then, the surrounding fluid serves as a loading device which can
do work on the solids. This changes the relevant thermodynamic potential from
F to G = F + pV, (8.2.28) sometimes called the Gibbs' function. Involved, tacitly,
is the assumption that the solids will stay in V, and not diffuse into the
surrounding fluid,
116 8. Mixture Theory 8.2 Elementary Theory 117 an assumption which can be
quite good, or very bad, depending on the materials involved. It is a poor
assumption for dry ice in air at room tem- perature, for example. Here, the
subvolumes are still considered to add up to V, as indicated by (8.2.1), but V can
vary. Also, (8.2.2) should hold with ME fixed. Again assuming non-zero
volumes and masses, we get, as equilibrium equations, Pl ..... Pm ---- P, (8.2.29)
as a replacement for (8.2.16) and, as before, (8.2.17), the equality of chem- ical
potentials. Here, V is not known but p is regarded as a given constant. Often,
workers will go a step further. Suppose p is a monotonically de- creasing
function of V1 when the other variables are held fixed (Opl/OV1 = -02F1/OV12
< 0). Then, we can solve p -- p for Vx = f(p,O, Me()), and express G = F + pV1
(8.2.30) as a function of these variables. Then, an exercise in calculus gives 0- --
Sl, Op -- V1, 0Me(1) - 13. (8.2.31) With similar assumptions, we can do the
same for the other ingredients and get m m G-- y.Gi-- y.(Fid-pVi) = F+pV, i= i=
(8.2.32) = G(O,p, Me(i)). Breakdowns in the assumed invcrtibility associated
with phase transitions sometimes occur. Attempts to perform the inversion can
then lead to Gibbs' functions which are multivalued and/or exhibit singularities
when the cor- responding Hclmholtz free energy function is well-behaved. This
is not to say that the latter cannot be ill-behaved, but the situation described is
rather common. Commonly, phase diagrams are drawn to indicate under what
conditions phase transformations take place. Typical diagrams for alloys are
discussed by Ricci [38]. After taking care of satisfying the equilibrium equations
one would like to test for stability and a common procedure is to try to use the
second derivative test. If, say, one is using (8.2.9) and is not concerned with the
possibility of zero masses or volumes, one would calculate the second dif-
ferential of EB, a quadratic in the differentials of Me (i) , Vi and $i. Some of the
differentials can be eliminated, expressed as linear combinations of others, to get
a quadratic form involving differentials which can be varied arbitrarily. For
metastability one would like this to be positive or maybe zero, for all choices of
the differentials. Schematically, one has a quadratic condition like Q = aijxixj _
0 (i,j=l) for all choices of xi, and we can assume that (8.2.33) aij - aji. If this is
not true, we can replace the cocffcients aij by (aij d- aji)/2, which does not affect
Q. One way of proceeding is to determine all of the cigcnvalucs of the matrix ]]
aij ]]; they must be nonnegative. In most cases it is easier to use another test.
Suppose first that the inequality is to be satisfied in the strict sense, Q = 0 only
when all xi = 0. Then, in the matrix ]] aij [[, go down the main diagonal,
blocking out i x 1, 2 x 2,..., matrices as indicated by the diagram [i, a13 a32 a33 [
aln aln a2n a3n ann Then, we will have Q > 0 if, and only if, the determinants of
all these matrices are positive? This gives a string of inequalities such as all > O,
alla22 -- a122 > 0,... , (8.2.34) not so difficult to check. If (8.2.33) does not hold
in the strict sense, there will be some nonzero xi for which Q -- 0 and, with Q _
0, these will satisfy the first derivative test for a minimum, aijxj ----- O, (8.2.35)
j=l linear equations which are fairly easy to solve. One can then make a change
of variables to get a quadratic form in fewer variables, which is strictly positive,
then use the above criterion for it. This covers some of the basic ideas used in
formulating elementary prob- lems involving mixtures, enough to indicate that
they are similar to ideas we have used before. As might be expected from our
discussion, analysis of physical problems of this kind tends to be rather
complicated but involves ideas much like those used before. 2A proof is given in
Frazer et al. [39].
120 8. Mixture Theory Now consider what happens if we let M approach zero,
which will also force m to approach zero. Since we are dealing with a solid, V is
not likely to approach V if we make V large enough, so the pressure should
approach zero. Under these conditions, if we vary 0, the pressure will remain
zero and the volume V1 can be expected to change. This is the analogue of the
thermal expansion discussed in Section 2.5. This gives a function which can be
measured V: v(O), when M = m = O, (8.3.11) such that the derivative in (8.3.8)
vanishes. With this, and (8.2.13), we then have dF1 = -1 dO, (8.3.12) where = S
[ = (O) (8.3.13) v =v(0) is the entropy function for the special processes
considered. To determine it, one needs measurements of the specific heat at zero
stress, the analogue of the Co(O) mentioned in Section 2.5. This will determine
to within an unimportant additive constant. Now, we can introduce a kind of
analogue of the strain energy func- tion introduced in Section 2.5. Bear in mind
that M is considered as a fixed constant, so the empirical functions could depend
on M were we to consider other values. Here, we introduce a function such that
W = W(V, m, 0), (8.3.14) w(v(o), o, o) = o, (8.3.15) OW OF1 OW OF1 OV1
OVi' (8.3.16) where the right sides of (8.3.16) are expressed as functions of Vx,
m and 0 using (8.3.7), (8.3.8) and (8.3.9). The idea is to integrate these
equations, subject to (8.3.15), to obtain W. Given suitable isothermal mechanical
data, this is enough to determine W uniquely. By arguments similar to those
used in Section 2.5, it then follows from above that (8.3.17) F 1 = W-/l(O)dO.
We will not elaborate on these matters. However, some things are worth noting.
There are the constants a and b which must be regarded as rather arbitrary,
physically. As is clear from (8.3.7), they do affect the values of 8.3 A Solid in an
Ideal Gas 121 chemical potentials, so these are not really so uniquely defined for
a given pair of materials. In turn, this affects F and, through it, the dependence of
the entropy $1 on m. For common kinds of stability analyses, such ambi- guities
seem not to matter. In other kinds of thermodynamic systems, am- biguities of
this kind become, in some sense, greater. As a second point, we note that, given
sufficiently smooth data, we could calculate 02F1/OmOV2 in two different
ways, either by differentiating (8.3.8) with respect to V, or by differentiating
(8.3.9) with respect to m. Equating the two gives o ( v- o (8.a.18) 0--11 In .' L ] =
mm . V - V ]' Sometimes, relations of this kind are used as a crutch, to try to get
the best estimates of the empirical functions from data which might be too sparse
or inaccurate to make a very reliable direct determination. Of course, one cannot
exclude the possibility that any theory may be overturned by contradictory
experimental results, but with rather old, generally accepted theories such as we
are considering, I think workers would be more likely to believe that it is the
experiment which is in error. Some interesting observations are reported by Gent
and Tompkins [40]. Briefly, a piece of rubber is placed in a pressure chamber
which is then filled with a gas. One then waits for some time to let the system
come to equilibrium, or at least very dose to this. Then the chamber is vented to
permit quick escape of the gas. If the pressure was high enough, the whole block
explodes, becoming a collection of small fragments. The explanation they
propose is as follows. Inevitably, a piece of rubber contains tiny holes. In the
first phase, the gas moves rather slowly into the rubber, hence into the holes.
Generally, diffusive motions like this are not very fast. As sug- gested by the
above analysis, the holes then fill with gas, at a pressure comparable to that in
the gas external to the sample. When the exterior gas is vented, that in the holes
cannot move out so quickly, so one still has the high pressure there. Relatively,
this is like imposing a high ten- sion on the solid and, intuitively, this could pull
it apart. Using this idea, combined with some rough calculations involving
rubber elasticity theory, they get an order of magnitude estimate of the value of
the pressure which must be exceeded to produce this kind of failure and it is
compatible with the observed values. As far as I know, no one else has proposed
a better analysis of the phenomenon. Phenomena somewhat like this are
sometimes explained in a very different way. If one ignores the possibility that
the gas enters the solid, one could argue that the sudden release of the exterior
pressure initiates something similar to a tensile shock or stress wave mov- ing
into the solid. Generally, such analyses use linear theory, with some ad hoc
estimate of how large the tension must be to produce breakage, and, usually, this
kind of argument predicts breakage at particular places, producing spalling of
plates, etc. It is not inconceivable that one could use some analysis of this
general kind to explain the phenomenon. Generally,
122 8. Mixture Theory one can encounter cases where two quite different
theories seem to explain the same phenomenon. Then, one tries to decide which
is best. Decisions of this kind are made, occasionally, but no simple rule book
describes how they are made. However, arguments about such matters can result
in the design of interesting experiments in order to test the merits of different
proposals. 8.4 8,2, Exercises For the thermodynamic experiments discussed in
Section 8.3, one assumption is inappropriate for a solid like dry ice, which
readily transfers some of its mass to air. Revise the analysis to apply such solids.
To simplify considerations, assume that air does not transfer any of its mass to
the solid. You may introduce one other simplifying assumption which seems
physically reasonable to you, based on your experience with such solids, after
clearly stating what it is. Discuss how the kinds of measurements needed differ
from those discussed before. In the three-dimensional theory of linear elasticity,
W, the strain en- ergy (per unit volume) function, is assumed to be an
homogeneous quadratic function of six measures of strain, and that it is positive
definite. For a transversely isotropic material, symmetry considera- tions reduce
this to the form W = Cl(x I -]- x2) 2 - C2x - C3(x I - x2)x3 - C4(x _]_ x5 2) '-
C5(x6 2 - XlX2) , where the C's are material constants and the x's label the
strains. What inequalities must the constants satisfy, for the quadratic to be
positive, in the strict sense? 9 Equilibrium of Liquid Crystals and Rods 9.1
Liquid Crystal Energies Although they are liquids, nematic liquid crystals
commonly used in display devices involve equilibrium problems which are more
like some encountered in solids. Thus, for example, one finds a chapter on them
in a volume on elasticity theory in the Landau-Lifshitz series on theoretical
physics [41]. At the same time, consideration of them will involve issues which
are different from those encountered before. Here, orientation replaces
deformation as an important quantity. For equilibrium situations, forces can be
involved, but they are generally not of sufficient interest to induce workers to try
to measure them and, more often than not, workers solve problems with- out
considering them explicitly. Also, for the first time, we will need to consider
effects associated with electromagnetic fields. Additionally, in the design of
devices and in measurement of material moduli, transitions play an important
role. These are liquids which are optically anisotropic when they are at rest,
being of what is often called the uniaxial kind, as are some crystals like Iceland
Spar. Roughly, such materials have one preferred direction, any direction
perpendicular to this being physically indistinguishable from any other. Suffice
it to say that one can use optical observations to determine the preferred
direction. We represent it by a unit vector n, n-n = 1, (9.1.1)
136 9. Equilibrium of LiquidCrystals and Rods Consider the solution for 0 _< z
_< L/2 where X > 0. Then (9.3.27) and (9.3.28) become d�> (9.3.29) kz = X/1 -
sin 2 0 sin 2 99 and kL (93.30) 2 J0 V/1 -- sin2 0 sin 2 qo Given k and L, the
idea is to solve (9.3.30) for 0. Sutituting this in (9.3.29) gives z a function of X,
the inverse of the function X(z) for 0 < z < L/2. To obtain the solution in the
upper half, put X(z) = x(L- z), L/2 < z < L, (9.3.31) it being a simple matter to
verify that this gives a solution of (9.3.27) satisfying (9.3.28). Concerning h(0),
given by (9.3.30) for 0 0 < /2, one can show that (a) h(0) is a monotonically
increasing, continuous function of 0; (b) lim h(o) = orr/2 (c) h(0) = This means
that kL determines a unique value of 0 in the range of interest provided kL eL : >
r. (9.3.32) From (9.3.19), this means that these solutions occur whenever
conditions are such that = 0 is unstable. For these nontrivial solutions, it follows
from (a) and (c) that as kL approaches the critical condition indicated by
(9.3.20), 0 approaches zero, so near this, it and hence sin 2 0, will be small. Then
(9.3.13) implies that will be small, close to zero. So, the distortion develops
rather smoothly as one increases e, passing through ec, again like the pitchfork
bifurcations we studied before. Actually, this makes it a little difficult to observe
exactly when the transition occurs: other methods can detect it, before it is
optically discernible. As one increases e, or, if you like, the somewhat fictional
voltage eL, kL increases so, from (a), 0 increases, n becoming more nearly
parallel to E near the mid-plane. The experience is that it only takes a few volts
to get n fairly close to the direction of E in a large fraction of the gap. In the
small parts near the walls, n then varies rapidly with position to adjust to the wall
orientation. Transitions somewhat similar to this occur in various other situations
of physical interest. To recall what we did, we recognized that, with an electric
9.4 Elastica Theory 137 (or magnetic) field, we could have unstable equilibria,
with n perpendicular to the field. We then introduced something else, the wall
effects, to help stabilize this. As we have seen, this can be successful if the field
is not too strong. If it is strong enough it gives a transition. Workers refer to the
transitions of this kind as FriedeWcsz transitions in honor of a Russian scientist
who discovered them experimentally using magnetic fields. The classical
methods for measuring K and Ka employ rather similar set-ups and transitions.
For K1, we start with the sample oriented as be- fore, but apply a field in the
direction normal to the plates, one case where the voltage across the plates is the
product of the field strength and the gap width. For Ka, the plates are treated to
make n align normal to them. The field is then applied perpendicular to this
direction. The corresponding cal- culations are discussed by various writers. For
example, they are included in three of the best general reference works on liquid
crystals, the books by Chandrasekhar [44], de Gennes [45], and Virga [46].
Display devices exploit Frederlcsz transitions, using fields strong enough to
induce n to become nearly parallel to the field except very near the walls.
Chandrasekhar dis- cusses the twisted nematic cell, which is used in many such
devices. The mathematical theory of nematic liquid crystals has undergone a
period of rapid development, as is discussed by Virga [46]. There is a point
worth bearing in mind. We considered only some very special types of
configurations. Those regarded as stable may turn out to be unstable with respect
to more general variations. Cohen [47] explored this theoretically for the set-up
used to measure K. He finds that, for many values of the moduli, there is an
instability which sets in at a field strength lower than that given by one-
dimensional theory, causing n to be nonuniform in a periodic manner.
Conversation with experimentalists indi- cates that some had noticed a little
undulation which could be associated with this, not so small in recently
discovered liquid crystals where K is relatively large, so there is some reason to
be concerned about this. So far at least, it has not been feasible to analyze the
sample of finite size which must be used in the experiments. So, as is typical,
one gets into more so- phisticated problems as one tries to take care of the
inevitable loose ends, in order to understand nature a little better. 9.4 Elastica
Theory It is a matter of common experience that the straight long thin bodies we
call rods, bars, beams and wires tend to depart from their straight shape, to
buckle when we apply compressive loads of moderate size at their ends.
Obviously, the bar theory discussed earlier cannot cope with this. One can use
somewhat similar one-dimensional theory, but it must let the line segments
become curves. The simplest possibility is that the curve is contained in some
plane, and, for a number of physically interesting
138 9. Equilibrium of LiquidCrystals and Rods situations, this is the case. The
simplest theory of this kind is Euler's theory of the Elastica, created by this
famous scientist in the eighteenth century. It has proved to be very good, so it is
still used by those concerned with designing safe structures. It may seem strange
to group these rods with liquid crystals. The reason is that, mathematically, the
more elementary analyses of Frederlcsz transitions are almost the same as the
Euler buckling problems. For this reason, functions like the h(q0) occurring in
(9.3.30), one of the so-called elliptic integrals is, much like the trigonometric
functions, long considered to be important enough to have their values listed in
tables of functions, now readily calculated by computer. The notion of a
reference configuration can be carried over unchanged from bars: the stable
unstressed configurations are considered as straight, either way. However, we
prefer to use different notation. Instead of x, we will write $, with 0 < $ < �
(9.4.1) describing the reference. A deformation will take this to some plane
curve, so we introduce rectangular Cartesian coordinates (x, y) in the plane. The
plane curve can then be represented parametrically by equations of the form x =
x(S), y -- y(S). (9.4.2) We can also introduce an analogue of the y(x) used in rod
theory; it is the arc length s, the distance measured along the curve, given by 4 s
= V/(x,)2 + (y,)2 dS = s(S). (9.4.3) In terms of this, we can introduce the stretch
A = s , (9.4.4) describing changes in length in the same way it did in bar theory.
In buck- ling phenomena, the bending deformations are generally large enough
to be perceived by the eye, but the changes of length tend to be relatively small.
The classical theory introduces an approximation associated with this which
does simplify the theory. It is that the rod is considered to be incxtcnsible,
incapable of changing its length, so, always, s = S, A = 1 = x 2 -t- y2 = 1. (9.4.5)
This makes the vector (x , y), tangent to the curve, a unit vector, repre- sentable
in the form x' = cos u, y' = sin u, (9.4.6) 4Readers not familiar with this can find
a discussion of it, curvature, etc., in any elementary book on differential
geometry. 9.4 Elastica Theory 139 u being the angle it makes with the x-axis. To
cover the fact that a rod offers some resistance to bending, a bending energy per
unit length W is introduced, a kind of strain energy function. For the ancient
theory, the assumption is that it is of the form 1 K(u,)2 ' w = w(o, = (9.4.7)
where K is a positive function of O called the flexural rigidity. From differen-
tial geometry y is the curvature of the curve. Different kinds of boundary
conditions, etc., are used to model different kinds of physical problems. Often,
but not always, they are such that the thermodynamic ideas of equilibrium are
applicable. We will consider one case where they are, as an illustrative example.
Suppose a rod is set into concrete, forming part of a floor, with the rod normal to
the floor. Choose axes so that the floor is the plane x -- 0, the rod's reference
configuration being on the x-axis. We will regard one end of the rod as being at
the origin, ignoring the piece buried in the concrete except for estimating the
boundary conditions. Physically, setting it in will fix the end position and also
fix the tangent vector at this end, say $ = 0, giving the boundary conditions x(0)
= y(0) = .(0) = 0. (9.4.8) Physically, no work will be done on this end. The sides
are to be left free so no work will be done here. As the words suggest, we
consider the system to be in the earth's gravitational field, giving a gravitational
force acting in the direction of the negative x-axis. We ignore the effect of this
on the rod itself. However, we will use it to dead-load the end $ = �. That is, we
will here firmly attach a rigid body with weight w. Associated with this is a
potential energy which can be taken as wxc, (9.4.9) where xc is the x-coordinate
of the centre of mass of this body. For sim- plicity, we assume it attached so that
this point is also the end of the rod, so xc = x(�). (9.4.10) So, overall, the
loading is conservative and, as usual, we assume 0 = const., one of the standard
situations covered by equilibrium theory. The applicable thermodynamic
potential can be taken as = K(u') dS + wx(�). (9.4.11) Here, there is a
possibility of having end-point minima, with the weight and perhaps part of the
rod, coming into contact with the floor. Clearly, this
140 9. Equilibrium of LiquidCrystals and Rods depends on the size and shape of
the weight. We will ignore this constraint and, for this reason, our analysis will
be incomplete. In part, the boundary conditions (9.4.8) can be taken care of, by
using (9.4.6) to write x(S) = cosydS, y(S) = sin vdS. (9.4.12) With this, we can
rewrite (9.4.11) in the form = /oe I K ( v')2 -k w cos vl dS, (9.4.13) with v still
subject to the boundary condition v(0) = 0. To determine the conditions needed
for equilibrium, we can, as usual, calculate 5 and equate it to zero. In the interior
of the interval, we get the equation which one would expect from (9.3.10), Kv" +
w sin u = 0, and the first integral suggested by (9.3.11). In addition, the analogue
of the last term in (9.3.9) produces another boundary condition: we should have
v(0) = v'(�) = 0. (9.4.14) Essentially, this is the liquid crystal problem we
analysed earlier. To see this, put v = = �/2, (9.4.15) so we have � (0) = o, = o,
(9.4.16) fitting conditions listed in (9.3.2) and (9.3.22). Also, (9.4.13) becomes
L/2 = [2K(') 2 + w(1 - 2 sin 2 r)] dS J0 (9.4.17) z,/2 = 2 [K(t) 2 - w sin 2 ] dS +
const., J0 the integrand being of the same form as in (9.3.16) with the constants
labeled in a different way. The upper limits look different so, in effect, we are
looking at the liquid crystal energy for the lower half of the sample. However,
from symmetry, it is easy to see that the upper half has the same energy, so one
can replace the upper limit in (9.3.16) by L/2 and multiply the new integral by 2.
Thus, with some bookkeeping, one can use the whole analysis for these two
problems, which are physically quite different. The critical condition
corresponding to (9.3.20) works out to be wc� 2 = Kr 2. (9.4.18) 9.5 Exercises
141 Thus, for w < wc the rod stays as is in its reference configuration. For w a
little larger than Wc, it will start to bend over a little. As the load increases,
v(�) will get closer to w, the rod bending nearly double. Realistically, some-
where along the line it will bend enough to let the weight contact the floor.
Consideration of this would, of course, make the two problems different. In this
case, it is not customary to try to use (9.4.18) to obtain an ex- perimental value
for K, partly because small misalignments etc. introduce noticeable errors in
such measurements. A classical estimate which seems good gives K -- El, where
E is Young's modulus, obtainable from a simple tension test and I is a
geometrical factor, a certain moment of inertia. One can get it directly using a
simple bending experiment, for example. These and some other aspects of
classical rod theory are discussed in more detail by Love [48], for example. In
previous studies we have concentrated more on issues relating to at- tempts to
find equations to describe response of materials. Here, this issue is fairly well
settled and we move into the province of the engineer who needs to resolve
stability questions in order to design safe structures. Such workers use various
ideas which have not been mentioned in our discus- sions, as can be seen from
the books of Leipholz [49] or Thompson and Hunt [25], for example. Numerous
other kinds of workers are interested in aspects of stability, producing a literature
so large as to be indigestible for one person. 9.5 9.2, Exercises An unusual
stability problem arises in liquid crystals, associated with what is called the
"plage tordue." The sample is prepared like that discussed in Section 9.3 except
that no field is applied. Instead, one plate is simply rotated relative to the other,
keeping them parallel, so the orientation at the two plates becomes different. To
analyze this, use (9.3.3) to calculate the energy. Then determine the equilib-
rium equation and boundary conditions, and find the solutions. You should get
two solutions for each angle of rotation, if you read Section 9.1 carefully.
Determine which of these is most stable and how this changes as the angle of
rotation increases. An experiment long used to determine values of K in (9.1.11)
em- ploys a sample of the kind described in Section 9.1, except that the field is
here normal to the plates. Instead of (9.3.3), assume that n = (cos X, sin X), X =
X(z). Substituting this in (9.3.3) gives 2W = (K1 cos 2 X + K3 sin 2 X)(X) 2, X
= X(z),
162 10. Reconsideration of Generalities for the time scales involved in the
observations. Still, it is hard to avoid the impression that equilibrium theory is
being misused but with some success. Some workers like to think in terms of a
"rubber plateau," involving time scales neither too short nor too long, over which
relaxation processes slow down enough to obtain something that looks much like
equilibrium, in an experiment designed to produce a static response. For
example, one view is that if one clamped the edges of a sheet to produce unequal
stretches, one could produce results similar to those discussed in Chapter 6,
involving some shear stresses, for a time which may seem very long. However,
the notion is that those shear stresses will eventually relax to zero. To see this,
one might need to observe them for times better measured in months or years.
Think of silly putty and you see behavior something like this but speeded up
considerably. I do not regard this as an unreasonable point of view. It does
suggest that true equilibrium theory might be that appropriate for a fluid and we
are using something quite different from this for the nonequilibrium processes
occurring on the plateau. Intuitively, we are then not close to real equilibrium.
Put one way, we seem able to use Gibbs' ideas to find some processes which, for
a limited time only, serve as attractors. Given this, there is some reason to
believe that it should be feasible to use more accurate viscoelasticity theory to
account for the slow changes and to associate with this a means of calculating
energy and entropy. Reasonably, this should not differ much from what we use
for rubber elasticity on the somewhat vaguely defined rubber plateau. There are
others who prefer to believe that, instead of reaching such a plateau, we are
getting close to real equilibrium, better described by theory more like the rubber
elasticity we have used. Also, there is a range of opinions about whether it is
feasible to make sense of entropy as it pertains to viscoelasticity theory and, if
so, how it should be done. The situation is not so clear that reasonable persons
cannot disagree about such matters, as I see it. What is involved is a kind of
theory too complicated to cover in an elementary way. However, I will try to
give some indication as to why such theories involve some unusual difficulties.
More often than not, nonlinear theory is needed to describe phenomena of
interest. However, linear one-dimensional theory is much simpler and is useful
for describing some phenomena, for example, small vibrations. Even here, the
basic difficulty appears. Often, the interest is in behavior in shear, so we will
consider the interpretation in terms of the shearing of plates. The difficulties are
more associated with the mechanics and there is some ques- tion concerning the
thermodynamics, so I will consider purely mechanical theory. When I first came
into contact with such work, it was common for workers to draw pictures of
springs and dashpots linked together in some fashion, using this to motivate
considering equations of the form M 0rat N y. am Ot m ---- y. bn On (10.5.2)
0tn, m=0 10.5 States 163 as a substitute for a constitutivc equation for r. Here, as
before, r and 3' are, respectively, the shear stress and strain. This is to be used in
conjunction with the equation of motion (2.4.5). The a's and b's are constant,
related to spring constants and dashpot viscosities in the picture, from which you
can infer their algebraic signs. Typically, M and N differ by one at most. If one
tries to match data on small vibrations, the experience is that, if you try to get by
with small values of M and N, you can only cover a very small range of
frequencies with resonable accuracy. As you try to increase the range of
frequencies covered, you need to increase M and N and there seems to be no
upper limit to this. In this respect, such models are not very satisfactory.
However, workers do still use them or nonlinear generalizations of them, to
analyze some kinds of phenomena. Setting all derivatives equal to zero, one gets
r proportional to 3'- This gives for the static elastic shear modulus, the value
bo/ao. If you conceive the material to be a fluid, set b0 = 0. Otherwise, any
difference between fluids and solids is more quantitative, more associated with
values of the constants estimated from experiment. In another respect, such
models are at least awkward. Typical relaxation experiments involve suddenly
imposing a deformation or force and holding it. So the time derivatives involved
are, initially, very large, perhaps infinite. With such experiences, workers were
encouraged to consider a differ- ent model proposed in 1874 by Boltzmann [65].
It replaces (10.5.2) by an equation of the form t): t) + n(t - t) -- (2O (10.5.3)
where /z is the equilibrium shear modulus and h(t - a) is a function depending on
the material. It is considered to decrease quite rapidly as cr -* -cx. In the sense
indicated, the material remembers all deformations occurring in the past but not
as well those that occurred a long time ago. Philosophically, it is not so pleasant
to think that one must know these ancient deformations but many workers find
ways to live with the idea. Also, there are the obvious practical difficulties in
knowing what to assume about this. If you prefer, you can think of starting with
having the plate undeformed ('7 = 0) up to a certain time, which alleviates the
problem but does not completely eliminate it. Some relevant experimentation is
covered by Ferry [66]. Three-dimensional linear theory is discussed by Leitman
and Fisher [67], for example. A stress relaxation experiment involves starting
with a sample which seems to be in equilibrium, with 3' = 0 for t < 0, suddenly
imposing a constant shear strain % and holding it, then measuring r as a function
of time. The aim is to obtain an experimental determination of I(t) = r(t)/7o.
(10.5.4)
170 10. Reconsideration of Generalities where I have used the first law. To
decide for yourself whether such a statement makes sense, a first step is to check
whether it does hold for systems familiar to you. If so, you can try to design and
build a system which violates it. If you cannot, you acquire some faith that the
law is sound, but there is no way to prove this. Let us try one check using the
thermoelastic theory of bars discussed in Chapter 2. For a bar, consider a cycle
beginning at t -- 0, ending at t = T, conform- ing to (10.6.4), so that 5Q = Qdt = -
Pdt = -AW _ 0. (10.6.5) To bring in the contact with the heat baths, we consider
the ends to be insulated, so q(L, t) -- q(0, t) = 0 D t, (10.6.6) and use Newton's
law of cooling, with the assumptions that, for 0 < x < L, r = c[O -- g(x,t)], 0 _ t <
tl, r = 0, tl _ t < r: 0[02 -- O(x, t)], t2 < t < ta, r = 0, ta < t < T, (10.6.7) with c
some positive constant, 0(x, t) being the temperature of the bar. It is not
immediately obvious that any such cycles exist, since some equations must be
satisfied, and I will not deal with this question. The question is more to check
whether any that may exist do conform to this version of the second law. To be
definite, by a cycle, we mean a process defined for times in the interval [0, T],
with y(x,T) = y(x,O) A(x,T) = A(x,O),O(x,T) : 0(x,0) (10.6.8) satisfying the
equation of motion (2.3.8) with f = 0 and the heat equation (2.3.27) with r as
prescribed in (10.6.7). One can allow for such things as the shock waves
discussed in Section 7.1 but I will leave it to you to determine whether this
affects the conclusions we will make. Now, with (10.6.8) we will have, in
particular L O) dx $(T) - $(0) = / q(A, = 0. (10.6.9) Then, from the global form
of the Clausius Duhem inequality (2.3.13), with Xl = 0, x2 = L, we obtain i T
foL(r/O) dxdt < O, (10.6.10) 10.6 Laws of Classical Thermodynamics 171 where
we have used (10.6.6). With (10.6.7), this gives ct [(0 - 0)/0] dt + [(02 - 0)/0] dt
dx <_ O. (10.6.11) From (10.6.4), we must also have a (0 - O) dt + (02 - 0) dt dx
> O. (10.6.12) Here the factor c, being positive, can be canceled. Now, with our
interpre- tation of this version of the second law, it should be impossible for heat
to be transferred from the colder to the hotter bath under the conditions assumed.
That is, briefly, Clausius : (02 O) dtdx > 0. (10.6.13) Odo Jr2 The question is
whether this really does follow from (10.6.10) and (10.6.12), perhaps by using
something else deducible from the bar theory considered. For this, it is helpful to
note that, for any possible values of 01 and 0, (01 - 0)(0 1 - 0 -1) 0 =: (01 - 0)/01
_ (01 - 0)/0. (10.6.14) Of course, one can here replace 01 by 02. Thus, using
(10.6.11), we have { [(o1- o) /o1] ctt + f i3[(o2 - o)/o2] ctt } ctx (10.6.15) <_
[(01 - 0)/0] at + [(02 - 0)/0] at ax _< 0. Now, use (10.6.12) to eliminate the first
term on the left to give - oi (02- o)ax at _< o, (10.6.16) and, with 02 > 0, this
does give the inequality alleged to hold in (10.6.13). So, as interpreted here, we
have verified that (10.6.3) does hold in the cases considered. It is not hard to
check that we would have encountered a violation had we considered ct to be
negative. Said differently, we have come close to proving that a ) 0 follows from
(10.6.3). To complete the proof, one would need to show that there is at least one
cycle of the kind considered. If you are willing to assume this, you have another
reason to think that (10.6.3) makes sense as interpreted here. Suffice it to say
that this old law is still accepted because it has been tested in numerous ways
and found to be trustworthy.
174 10. Reconsideration of Generalities Read on, and you will see that Serrin
regards this as a very different version, not equivalent to the previous two. Two
things are worth noting. It is more than glossing over points of rigor to fail to
mention that those states should be describable by a finite number of parameters,
as Pippard does. Also, what is a state a0 in (10.6.20) is, in Pippard's version, an
"equilibrium state." So, accept that the world of classical thermodynamics is
limited to the cases which can be described in terms of states which are, in turn,
describable by a finite list of parameters. Further, such states are to fit the
description of equilibrium states discussed earlier. From what else he does, this
is what I infer to be Pippard's view of what is classical thermodynamics and, in
this respect, he has lots of company. Workers who accept this view generally
accept the idea that energy and entropy are well-defined as functions of such
states. This view is too limited for some, Bridgman for one, and I have declared
myself to be with him in this. I have said enough about the problems met in
deciding what should be meant by states to indicate that, sometimes, I do not see
how to use the third version in cases where the other two can be applied. If I had
some conjecture about this, I would return to the previous versions to try to
decide whether it made sense. In this respect, those versions are certainly not
replaceable by the third version for my purposes. The third version has some
merit in providing a basis for simple equilibrium studies such as we have
performed, when one does not want or know how to place these in the context of
some nonequilibrium theory. However, even here, it is hard to see how one
could use this to motivate or justify what we took to b.e basic definitions of
equilibrium. Neither does it seem to warn us against trying to use them for
problems involving sliding friction, for example. As I see it, the view adds little
to what Gibbs said about equilibrium. For such reasons, I do not find the third
version very useful, but you might. Easier to use, when it applies, is the Clausius
Duhcm inequality, which really comes in two versions. We have already
discussed this in relation to one-dimensional theories. For three-dimensional
continuum theories, the more traditional version goes as follows. Some fixed set
of material is con- sidered to be the thermodynamic system. Here, entropy is
introduced from the start. Introduce a reference configuration, some possible
configuration as a three-dimensional analogue of what we have used for bars and
plates. Let denote the region it occupies. With $ denoting the entropy, the
entropy per unit reference volume, we have, by assumption S = f rldv, (10.6.21)
dv being the volume element. Here, to describe Q, we introduce a heat flux
vector q and set Q = fo q' da, (10.6.22) 10.6 Laws of Classical Thermodynamics
175 where 0 denotes the boundary of , da the element of area and qn the
projection of q on the outward normal to the surface. Then, the Clausins- Duhem
inequality is dS/dt > fo (q,/O) da, (10.6.23) where, as usual, 0 is the temperature
of the material which, in general, varies with position and time. By using the
divergence theorem, etc., one can obtain a corresponding inequality which holds
pointwise, but I will not belabor this. If you return to Chapter 1 and read the
quotation from Gibbs, you will find that rather puzzling integral, in which t
denotes ...the temperature of the part of the system receiving it .... It is not clear
what he had in mind but it seems not to be quite the integral occurring in the
Clausins inequality (10.6.17); something more like that occurring in (10.6.23)
seems to be the case. With the assumption that the quantity on the left vanishes
when integrated over a cycle, one gets an inequality somewhat like the Clausius
inequality, fti fO (q,/O)dadt < O. (10.6.24) In the above discussion of bars we
mentioned the possibility of putting different heat baths in contact with the ends
of a bar. Kestin [70] uses a similar idea to conclude that (10.6.23) is deducible
from our (1.2.7) when entropy is defined using the hypothesis of local
equilibrium. Possibly, one could use a similar argument to deduce that (10.6.24)
follows more generally from, say, Clausius' version of the second law. This
could be useful in trying to justify the assumed existence of entropy for more
general systems by this route, but I am not sure of this. The other version of this
inequality has been frequently used since it was suggested by Truesdell and
Toupin (p. 258 of [71]). For bars the quantity r may be thought of as being
defined within the bar. The proposal is to modify (10.6.22) in that analogous
way, writing Q=foq,da+frdv, (10.6.25) replacing (10.6.23) by dS/dr >_
fo(q,/O)da + fob(r/O)dv. (10.6.26) Then, (10.6.24) is modified in the obvious
way to ii [fo(q/O)da + /n(r/O)dv] dt < O. (10.6.27)
176 10. Reconsideration of Generalities It seems less likely that one could
produce a satisfactory derivation of this from any of the other three versions of
the second law. For such reasons, it is at least awkward to tackle problems
related to the existence of entropy with suitable properties, for some systems of
this kind, in particular theories of viscoelasticity. Motivated by such
considerations, Serrin [69] proposed a different version of the second law which
does seem to deliver useful results more easily. In spirit, it is not so different
from the first two versions and, for simpler kinds of theories, it leads to the same
conclusions. Like them, it deals with cycles. Consider any thermodynamic
system, letting C denote any cycle associated with it. Associate with this a
function depending on a parameter O, A(C, O) called the accumulation function,
with the interpretation that A(C, O) = heat received by the system during the
cycle, at temperatures 0 _ O. (10.6.28) Thus, for 5Q interpreted as in (10.6.1),
we have 5Q = A(C, ). (10.6.29) In words, his version is The accumulation
function of a nonadiabatic cycle process cannot be nonnegative. (10.6.30) He
explains how this relates to the Clausius and Kelvin Planck versions, as he
interprets them and, for this, I refer the reader to his paper. He shows that this
statement implies that O-2A(C, 6}) dO < 0. (10.6.31) Under assumptions which
are not very restrictive, one can put this in a form which looks more like the
Clausius inequality. Suppose that, for C fixed, A is a sufficiently smooth
function of O to enable us to integrate by parts. Suppose also that lim A/O = lim
A/O = 0. (10.6.32) From (10.6.29), the first condition will certainly be satisfied
if 5Q < , the second if no heat is received at temperatures near absolute zero and
even less restrictive assumptions would suffice. With these assumptions,
integration by parts gives, for any cycle, o 0 - dA < O. (10.6.33) Roughly, dA
represents the heat received by the system between the tem- peratures O and
O+dO, making this somewhat like the Clausius inequality. 10.6 Laws of
Classical Thermodynamics 177 Unlike the latter, there is no stipulation on how
heat baths, etc., are to be used to effect heat transfer. This also resembles the
rather mysterious inte- gral mentioned by Gibbs, providing one interpretation
which is meaningful. Let us try evaluating (10.6.33) for a thermoelastic bar. For
simplicity we will ignore heat conduction, so that L Q = fo rdx. (10.6.34)
Consider a cycle beginning at t = 0 and ending at t = T > 0, described by
functions y(x, t) and O(x, t), which should, of course, be defined for the whole
bar, 0 _ x _ L. The relevant points (x,t) in the x-t plane then form a rectangle R.
For r, one can use Newton's law of cooling, the Stefan Boltzmann law, or simply
the function which satisfies (2.3.12) when some inputs are used to calculate the
remaining terms. For simplicity, we assume that r(x,t) and O(x,t) are continuous
functions on the rectangle, which implies that they are bounded. Also, we
assume that O(x,t) is bounded away from zero, so, for some positive constants a
and b, we have 0 < a < O(x, t) _ b < cx:. (10.6.35) For convenience, we have
picked a to be smaller than the minimum value of O(x, t). Now divide up the
temperature interval into N equal parts, n = 1,..., N, giving intervals as indicated
by In : On < 0 <_ On+, O = a, ON+i = b, (10.6.36) On+ - On = (b- a)/N. Now,
we define subsets an of the rectangle R by the condition that (x, t) E an = O(x, t)
In. (10.6.37) No point has two temperatures associated with it, so one point
cannot belong to two such sets. Also, every point in R has a temperature
associated with it, so the union of these disjoint sets is R, N U an: R. (10.6.38)
Now, from (10.6.28) and (10.6.36), for the cycle C considered, A(C,O)=0 ifO<a,
(10.6.39) A(C,O)= rdxdt = rdxdt, ifO_b, which is consistent with (10.6.32). It is
then clear that dA = 0, except when O is in the interval a ( O _ b, so the left side
of (10.6.33) takes the form O 0 --1 dA. (10.6.40)
178 10. Reconsideration of Generalities Now, from at least one definition of the
integral we can approximate it as closely as we like in the manner indicated by N
jab) -1 dA - )i[A(C, )n+l) -- A(C, )n)]. (10.6.41) Now, the difference in the
square brackets is, from the definition of A, the heat supplied at temperature O in
the range On < O _< On+, (10.6.42) that supplied at 0 = On being counted in
both A's. From the definition of an, this is given by r dx dt. Thus, the above sum
can be written as n+ rdxdt. (10.6.43) Now, using (10.6.38), we also have N
/R(r/O)dxdt-- (r/O)dxdt, (10.6.44) and, with this, we can add and subtract this
term, to put (10.6.41) in the form N Note that, if N is large, 0 is nearly equal to
On+ in an. By a more detailed analysis, which I will omit, this sum can be made
arbitrarily small by taking N large enough. In the limit as N - x, we obtain 0 --1
dA = 0 - dA = r/O) dx dt < 0, (10.6.46) which is what we would get by applying
the Clausius Duhem inequality to a cycle when heat conduction is neglected (q --
0). With a similar, but more tedious analysis, one can account for heat
conduction which modifies (10.6.46) in the manner expected from analogous
calculations based on the Clausius-Duhem inequality. Ignored here is the
possibility that the sets an may be so complicated that integrals over them are not
well-defined. For a rigorous treatment, one needs analysis too complex to be
used here, 10.6 Laws of Classical Thermodynamics 179 involving the theory of
Lebesgue integration, but it is possible. In the three- dimensional case, when Q is
given by (10.6.25), one can use similar argu- ments to show that, formally,
(10.6.27) is equivalent to (10.6.33). For var- ious kinds of theories of solids and
fluids, the Clausius-Duhem inequality applies and gives sensible results. So this
really provides numerous checks on the soundness of this version of the second
law. With, say, plasticity theories, where one has doubts about the existence of
entropy, it would be unsafe to assume that the Clausius-Duhem inequality
applies. However, if you accept that (10.6.25) applies, you can use (10.6.27) as a
mathematical representation of the second law, if you also accept this version. At
least from my point of view, this version fits comfortably as part of the subject
of classical thermodynamics, although it is relatively new. I prefer it to any of
the three versions mentioned before. There are also some different versions of
the third law which deals with the unattainability of absolute zero (8 = 0). As
stated by Pippard [4], two of these are By no finite series of processes is the
absolute zero attainable, (10.6.47) and, what is perhaps better regarded as an
addendum, As the absolute temperature tends to zero, the magnitude of (10.6.48)
the entropy change in any reversible process tends to zero. This provides support
for the view that the second equality of (10.6.32) is not a very restrictive
assumption, among other things. Various writers discuss this law and I will leave
it to the interested reader to pursue this on his own. Finally, some slightly
different versions of another law, the zeroth law, are mentioned by various
authors. As I interpret this, it is more concerned with the theory of temperature
which I have not tried to discuss. In their treatment of this, Fosdick and
Rajagopal [72] take as basic a version due to Maxwell, (Maxwell) Bodies whose
temperatures are equal to that of (10.6.49) the same body have themselves equal
temperatures. In statements used by some others, the notion of equality of
temperature is replaced by a statement that, pairwise, the bodies are in
equilibrium with each other, a notion that I find misleading and confusing.
Newer than the theory of the old engines but certainly old enough to be regarded
as classical, is another branch of thermodynamics, more re- lated to statistical
molecular theory. In this, Gibbs was also a pioneer. This involves quite different
ideas about concepts of temperature, energy and entropy, more related to
statistical averages. That these names are used here reflects the fact that some
situations can be analysed with this and the older theory, with compatible results,
if one agrees to use these names
182 [lO] [11] [12] [13] [14] [15] [16] [17] [18] [19] References Pippard, A.B.
(1985), Response and Stability, Cambridge University Press, Cambridge.
Gelland, I.M., and Fomin, S.V. (1963), Calculus of Variations (trans. R.
Silverman), Prentice-Hall, Englewood Cliffs, NJ. Rivlin, R.S. (1975), The
thermomechanics of materials with fading memory, in Theoretical Rheology,
Applied Science Publishers Ltd, London. Bell, J.F. (1973), The experimental
foundations of solid mechanics. Handbuch der Physik, vol. Via/l, Springer-
Verlag, Berlin. Kahl, G.D. (1967), Generalization of the Maxwell criterion for
van der Waals' equation. Phys. Rev., 155, 78-80. Ericksen, J.L. (ed.) (1984),
Proceedings of Workshop on Orienting Polymers, Minneapolis 1983. In Lecture
Notes in Mathematics, vol. 1063, Springer-Verlag, Berlin. Landau, L.D. (1965),
On the theory of phase transitions, in Collected Papers of L.D. Landau (ed. D.
Ter Haar), Gordon and Breach and Pergamon Press, New York. Thomas, L.A.,
and Wooster, W.A. (1951), Piezocrescence: The growth of Dauphin twinning in
quartz under stress. Proc. Roy. Soc. London, A205, 43-62. Perkins, J. (ed.)
(1975), Shape Memory Effects in Alloys, Plenum Press, New York. Delay, L.,
and Chandrasekaran, L. (eds.) (1982), Proc. Int. Conf. on Martcnsitic
Transformations. J. Physique, 43. [20] Nishiyama, Z. (1978), Martensitic
Transformations, Academic Press, New York. [21] [22] [23] James, R.D., and
Kinderlehrer, D. (1989), Theory of diffusionless phase transitions, in Partial
Differential Equations and Continuum Models of Phase Transitions (eds. M.
Rascle, D. Serre, and M. Slem- rod), Lecture Notes in Physics, vol. 344,
Springer-Verlag, New York. Sengers, A.L., Hocken, R., and Sengers, J.V.
(1977), Critical point universality and fluids. Phys. Today, 30, 42-51. Beatty,
M.F. (1987), Topics of finite elasticity: hyperelasticity of rub- ber, elastomers,
and biological tissues. Appl. Mech. Rev., 40, 1699- 1734. References 183 [24]
Kitsche, W., Miiller, I., and Strehlow, P. (1987), Simulation of pseu- doelastic
behavior in a system of balloons, in Metastability and In- completely Posed
Problems (cds. S. Antman, J.L. Ericksen, D. Kinder- lehrer and I. Miillcr), IMA
Volumes in Mathematics and Its Applica- tions, vol. 3, Springer-Verlag, New
York. [25] Thompson, J.M.T., and Hunt, G.W. (1973), A General Theory of
Elastic Stability, John Wiley and Sons, New York. [26] Alexander, H. (1971),
Tensile instability of initially spherical balloons. Int. J. Eng. Sci., 9, 151-162.
[27] Trcloar, L.R.G. (1948), Stresses and birefringence in rubber subjected to
general homogeneous strain. Proc. Phys. Soc., 60, 135-144. [28] Kearsley, E.A.
(1986), Asymmetric stretching of a symmetrically loaded elastic sheet. Int. J.
Solids Structures, 22, 111-119. [29] Haughton, D.M., and Ogden, R.W. (1978),
On the incremental equa- tions in non-linear elasticity - II. Bifurcation of
pressurized spherical shells. J. Mech. Phys. Solids, 26, 111-138. [30] Chen, Y.-
C. (1987), Stability of homogeneous deformations of an in- compressible elastic
body under dead-load surface tractions. J. Elas- ticity, 17, 223-248. [31] Chen,
Y.-C. (1988), Stability of pure homogeneous deformations of an elastic plate
with fixed edges, Q. J. Mech. Appl. Math., 41, 249-264. [32] Treloar, L.R.G.
(1949), The Physics of Rubber Elasticity, Clarendon Press, Oxford. [33]
Courant, R., and Friedrichs, K.O. (1948), Supersonic Flow and Shock Waves,
Interscience, New York. [34] Dunn, J.E., and Fosdick, R.L. (1988), Steady,
structured shock waves. Part 1: Thermoelastic materials. Arch. Ration. Mech.
Anal., 104, 295 365. [35] Gurtin, M.E. (1988), Toward a nonequilibrium
thermodynamics of two-phase materials, Arch. Ration. Mech. Anal., 100, 275
312. [36] Cheng, B.W. (1981), Polymer Surfaces, Cambridge University Press,
Cambridge. [37] Wu, S. (1982), Polymer Interface and Adhesion, Dckker, New
York. [38] Ricci, J.E. (1966), The Phase Rule and Heterogeneous Equilibrium,
Dover, New York.
184 [39] References Frazcr, R.A., Duncan, W.J., and Collar, A.R. (1938),
Elementary Matrices, Cambridge University Press, Cambridge. [40] Gent, A.N.,
and Tompkins, D.A. (1969), Nucleation and growth of gas bubbles in
elastomers. J. Appl. Phys. 40, 2520-2525. [41] Landau, L.D., Lifshitz, E.M.,
Kosevitch, A.M., and Pitaevskii, L.P. (1986), Theory of Elasticity, (3rd ed.)
(trans. J.B. Sykes and W.H. Reid), Pergamon Press, Oxford. [42] Frank, F.C.
(1958), On the theory of liquid crystals, Disc. Faraday $oc., 23, 19 28. [43]
Courant, R., and Hilbert, D. (1953), Methods of Mathematical Physics, vol. 1,
Interscience, New York. [44] Chandrasekhar, S. (1977), Liquid Crystals,
Cambridge University Press, Cambridge. [45] de Gennes, P.G. (1974), The
Physics of Liquid Crystals, Clarendon Press, Oxford. [46] Virga, E.G. (1994),
Variational Theories of Liquid Crystals, Chap- man & Hall, London. [47]
Cohen, R.A. (1988), Fractional step methods for liquid crystal prob- lems, Ph.D.
Thesis, University of Minnesota, MN. [48] Love, A.E.H. (1944), A Treatise on
the Mathematical Theory of Elas- ticity, 4th ed., Dover, New York. [49]
Leipholz, H. (1970), Stability Theory, Academic Press, New York. [50]
Grindlay, J. (1970), An Introduction to the Phenomenological Theory of
Ferroelectricity, Pergamon Press, Oxford. [51] Brown, W.F., Jr. (1966),
Magnetoelastic interactions, in Springer Tracts in Natural Philosophy, vol. 9 (ed.
C. Truesdell), Springer- Verlag, New York. [52] van der Waals, J.D. (1895),
Th5orie thermodynamique de la cap- illaritd dans l'hypothse d'unc variation
continual de densitd. Arch. Neefl. $ci. Exactes Nat., 28, 121 209. [53] Dunn,
J.E., and Scrrin, J. (1985), On the thcrmomechanics of inter- stitial working.
Arch. Ration. Mech. Anal., $$, 95 133. [54] Coleman, B.D., and Noll, W.
(1963), The thermodynamics of elastic materials with heat conduction and
viscosity. Arch. Ration. Mech. Anal., 13, 16178. References 185 [55] Miiller, I.
(1973), Thermodynamik, Die Grundlagen der Material The- or/e, Bertelmann
Universit/tsverlag, Diisseldorf. [56] Liu, I.-S. (1972), Method for exploitation of
the entropy principle. Arch. Ration. Mech. Anal., 46, 131-148. [57] Man, C.-S.
(1985), Dynamic admissible states, negative absolute tem- perature and the
entropy maximum principle. Arch. Ration. Mech. Anal., 9, 263-289. [58]
Poincard, H. (1885), Sur l'dquilibre d'une masse fiuide anime d'un mouvement
de rotation. Acta Math., 7, 259 380. [59] Ericksen, J.L. (1992), Reversible and
nondissipative processes. Q. Jl. Mech. Appl. Math., 45, 545-554. [60] Meixner,
J. (1973), Consistency of the Onsager-Casimir reciprocal relations. Advance
Mol. Relaxation Processes, 5, 319 331. [61] Wollants, P., Roos, J.R., and
Delaey, L. (1980), On the stress- dependence of the latent heat of
transformations as related to the efficiency of a work performing cycle of a
memory engine. $cripta Met., 14, 1217-1223. [62] De Groot, S., and Mazur, P.
(1962), Nonequilibrium Thermodynam- ics, North-Holland, Amsterdam. [63]
Green, A.E., and Naghdi, P.M. (1955), A general theory of an elastic- plastic
continuum. Arch. Ration. Mech. Anal., 18, 251-281. [64] Chaboche, J.L. (1988),
Continuum damage mechanics. Parts I and II. J. Appl. Mech., 55, 59-72. [65]
Boltzmann, L. (1874), Zur Theorie der elastischen Nachwirkung, Sitzber.
Kaiserl. Akad. Wiss. Wien Math.-Naturw. Kl., ?0(II), 275- 306. [66] Ferry, J.D.
(1970), Viscoelastic Properties of Polymers (2nd ed.), Wiley-Interscience, New
York. [67] Leitman, M.J., and Fisher, G.M.C. (1973), The linear theory of vis-
coelasticity, in Handbuch der Physik, vol. Via/3 (ed. C. Truesdell), Springer-
Verlag, New York, pp. 1-123. [68] Coleman, B.D., and Owen, D.R. (1974), A
mathematical foundation for thermodynamics. Arch. Ration. Mech. Anal., 54, 1-
104. [69] Serrin, J. (1979), Conceptual analyses of the classical laws of thermo-
dynamics. Arch. Ration. Mech. Anal., 70, 355-371.
186 [70] References Kestin, J. (1990), A note on the relation between the
hypothesis of local equilibrium and the Clausius Duhem inequality. J. Non-
Equilibrium Thermodynamics, 15, 193 212. [71] Truesdell, C., and Toupin, R.A.
(1960), The classical field theories, in Handbuch der Physik, vol. III/3. (ed. S.
Flfigge), 226-793, Springer- Verlag, New York. [72] Fosdick, R.L., and
Rajagopal, K.R. (1983), On the existence of a manifold for temperature. Arch.
Ration. Mech. Anal., 81,317-332. [73] Zanzotto, G. (1988), Twinning in
minerals and metals: remarks on the comparison of a thermoelastic theory, with
some available ex- perimental results, notes I and II. Atti. Ace. Lincei Rend. Fis.,
82, 723-741 and 743 756. [74] Truesdell, C. (1984), Rational Thermodynamics
(2nd ed.), Springer- Verlag, New York. [75] ilhavy, M. (1997), The Mechanics
and Thermodynamics of Contin- uous Media, Springer-Verlag, New York. Index
Accumulation function, 176 Austenite, 57 Ballistic free energy, 7-8, 14, 38, 40,
112 Biaxial stretch, 79-83 see also Treloar instability Bifurcation diagrams, 62-
65 pitchfork, 63, 87-89, 134-135 sub-critical, 64 Black-body radiation, 144 Body
couples electric, 129 magnetic, 129 Breaking bars, see Discontinuities Brownian
motion, 7, 150 Buckling bar, 13141 see also Euler's elastica Carath6orodory's
second law, 173-174 Chemical potentials, 111, 114 Clausius-Clapeyron
equation, 49, 66 Clausius-Duhem inequality one-dimensional form, 12, 14, 17,
25, 93, 154 three-dimensional form, 174 175 Clausius' inequality, 172 Clausius-
Planck inequality, 4, 13 14, 155 Clausius' second law, 169-171 Concentrations,
115 Conical boundary conditions, 129-130 Convex functions, 16, 60, 82 Cycles,
156-157, 166-167, 168, 170, 176 Cyclic processes, see Processes, cyclic
Difficulties in applying thermodynamics, 7, 26, 34, 143-146, 161-166
Discontinuities in bars breaking, 99 in equilibria, 50, 52-54 shock waves, 93-99
in plates, 61-62 peeling, 100-106