THERMOMECHANICS OF PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
This page is intentionally left blank
Series on Advances in Mathematics for Applied Sciences - Vol. 13
THERMOMECHANICS OF PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Antonio Romano Dipartimento di Matematica e Applicazioni Universita degli Studi di Napoli "Federico II"
World Scientific Singapore • New Jersey • London • Hong Kong
Published by World Scientific Publishing Co. Pte. Ltd. P O Box 128, Farrer Road, Singapore 9128 USA office: Suite IB, 1060 Main Street, River Edge, NJ 07661 UK office: 73 Lynton Mead, Totteridge, London N20 8DH
Library of Congress Cataloging-in-Publication Data Romano, Antonio. Thermomechanics of phase transitions in classical field theory/ Antonio Romano. p. cm. -- (Series on advances in mathematics for applied sciences ; vol. 13.) Includes bibliographical references. ISBN 9810213980 1. Phase transformations (Statistical physics) 2. Surfaces (Physics) 3. Field theory (Physics) 4. Continuum mechanics. I. Title. II. Series. QC175.16.P5R65 1993 530.4'74-dc20 93-30499 CIP Copyright © 1993 by World Scientific Publishing Co. Pte. Ltd. All rights reserved. This book, or parts thereof, may not be reproduced in any form orby any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the Publisher. For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 27 Congress Street, Salem, MA 01970, USA.
Printed in Singapore by JBW Printers & Binders Pte. Ltd.
PREFACE
The
phase
transitions
in a substance
C are due
to certain
weak
interactions among the molecules of C which might correctly be described by Q u a n t u m Mechanics. However, if we abstain from understanding
every
minute detail of behaviours in phase transitions, we can limit ourselves to describe
these phenomena
by resorting to much simpler and
tractable
schemes. In other words, we use a macroscopic description which is unable to explain either why a state change takes place or which modifications this produces at a microscopic level. T h e first approach starts from the experimental observation that the different phases are separated by very narrow layers across which the fields associated to C vary continuously but sharply. Owing to the high values assumed
by
the
gradients
of the fields
in
these layers,
the
ordinary
constitutive equations are supposed to depend weakly on the higher order gradients of the aforesaid fields. This, in turn, implies that the highest order derivatives
in
the
local
equations
of balance
are multiplied
by
small
coefficients. As is well-known, this property leads just to boundary layers whose localization and forms depend on the equations as well as on the region occupied by C (see for instance [50,51]).
v
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
VI
Another approach to attain a macroscopic description of phase transi tions
is obtained by resorting to the models of continua with an interface.
T h e basic idea of this approach is t h a t we can substitute the narrow layers between the phases with surfaces of discontinuity for the volume fields provided t h a t we associate to these surfaces some physical attributes which evoke the complex structure of the fields in the layers they substitute. This essay is just dedicated to this approach without any a t t e m p t to compare it with the alternative approach of boundary layers (see for instance [52] for a comparison). After some kinematical preliminaries (chapters 1 and 2), the detailed analysis of the balance equations for a system with an interface as well as of the average procedures which allow to define the physical attributes to associate to the interface are discussed in chapter 3. The remaining chapters are dedicated to the applications of the general scheme to the equilibrium and
thermodynamical
evolution
of phase
transitions
in liquids,
solids,
mixtures, crystals and ferroelectric materials. For the phase transitions in these last two materials a suitable use of the nonlocal theories of C o n t i n u u m Mechanics is suggested.
CONTENTS PREFACE
v
Chapter 1: GEOMETRY OF SURFACES 1.1 Definitions of regular surfaces, holonomic bases and metrics
1
1.2 The second fundamental form and the Gauss-Weingarten equations
5
1.3 Normal curvature
9
1.4 Elliptic, hyperbolic and parabolic points. Mean and Gaussian curvatures
11
1.5 Lines of curvature
14
1.6 Surface gradient operator and Gauss theorem
16
Chapter 2: KINEMATICS OF SURFACES 2.1 Velocity of a moving surface
23
2.2 The Thomas derivative
28
2.3 The velocity of a moving curve on a moving surface
30
vii
Vlll
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
2.4 Time derivative of a volume integral
31
2.5 Time derivative of a surface integral
34
2.6 Two time derivation formulas derived from (2.39) and (2.44).... 38
Chapter 3: BALANCE LAWS FOR A CONTINUOUS SYSTEM WITH AN INTERFACE 3.1 General balance laws
42
3.2 Mechanical modelling of nonmaterial interfaces
48
3.3 Balance equations for a system with an interface
58
3.4 The reduced dissipation inequalities
65
3.5 Constitutive equations
67
Chapter 4: PHASE EQUILIBRIUM 4.1 Phase equilibrium boundary value problems
73
4.2 Equilibrium of fluid phases separated by a planar interface in the absence of body force
76
4.3 A brief resume of phenomenology of state changes
78
4.4 Equilibrium of fluid phases separated by nonplanar interfaces
84
CONTENTS
— — —
■—
ix
4.5 Equilibrium of fluid phases separated by spherical interfaces
86
4.6 Equilibrium between isotropic solid and fluid phases
93
4.7 Variational formulation of phase equilibrium
97
Chapter 5: STATIONARY AND TIME DEPENDENT PROBLEMS 5.1 A stationary problem and its nondimensional analysis
105
5.2 On the approximate evolution of solid-liquid state changes
112
5.3 On the approximate evolution of liquid-vapour state changes
119
5.4 The case of a perfect gas
124
Chapter 6: PHASE CHANGES IN MIXTURES 6.1 Balance laws in classical mixtures
131
6.2 Constitutive equations and dissipation inequalities
134
6.3 Phase equilibrium in a binary fluid mixture
140
6.4 The influence of mass adsorption on surface tension
143
6.5 The Gibbs principle for phase equilibrium in fluid mixtures
147
6.6 The evaporation of a fluid into a gas
149
X
PHASE TRANSITIONS IN CLASSICAL FIELD THEORIES
Chapter 7: CRYSTAL GROWTH 7.1 Gibbs' and Wulffs equilibrium laws
159
7.2 About Wulffs construction
166
7.3 An introduction to nonlocal thermomechanical theory
169
7.4 Nonlocal balance laws
173
7.5 Nonlocal reduced dissipation inequalities
181
7.6 Preliminary considerations on crystal growth
183
7.7 A mathematical model of crystal growth in a binary nonreacting mixture
187
7.8 On the crystal equilibrium
194
7.9 T h e energy equation and the reduced dissipation inequality
199
7.10 The free boundary value problem describing the crystal growth
204
Chapter 8: SYSTEMS WITH INTERFACES AND FERROELECTRICITY 8.1 A brief survey of ferroelectricity
208
8.2 Maxwell's equations for a moving continuous system
211
8.3 A system of a rigid ferroelectric crystal in the presence of conductors and quasi-static approximation
213
8.4 The balance of energy and the entropy inequality
221
8.5 Constitutive equations and thermodynamical restrictions
224
CONTENTS
xi
8.6 Structure of the Weiss domains
228
8.7 The boundary value problem for the Weiss domains
234
8.8 Weiss' domains in the absence of electric field at equilibrium
REFERENCES
236
245
Chapter 1
GEOMETRY OF SURFACES 1.1 Definitions of regular surfaces, holonomic bases and metrics.
Let § 3 be the Euclidian three-dimensional space and (O, et) , i = 1, 2, 3, —> an orthonormal frame of reference in 8 3 . We denote by r = OP = x'e^ the position vector with respect to (0,e t ) of any point P e 8 3 . Let $ be a regular surface of 8 3 and r = r(w\«2),
(1.1)
a local parametrization of if. The regularity hypothesis implies that the scalar functions
I^IV."2).
(L2)
are of class C 1 and the rank of their Jacobian matrix is 2. The relations _ _ _ _ dr _ dx' <*-,<*-dua~dua I
•'
(, K
o\ >
2
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
a = 1, 2, define two vectors which are tangent to the coordinate curves on $. These vectors are linearly independent at all points r e J since the previous hypotheses imply that
(1.4)
lxa2\^0. I\a<*1 X «2 I # ° ■
Therefore, the vectors aa form a basis for the tangent space of If which is called a
holonomic a
coordinates (u ).
or
coordinate
basis associated
with the
curvilinear
From (1.1) and (1.3) we derive the square of the line
element which connects the points r and r + dr of $
ds2 = (dr)2 = aa/3 d^du13,
(1.5)
where ^ ^ a, V ^ - ^l J ^ f=^ aa, = aa.a,=
a
a
(1.6) (1-6)
are the metric coefficients. Since *■*'
\a1xa2\
=\det{aafi)
,
it follows that
v
' It is sufficient to observe that \ax X a 2 | = ^ ( a j X a 2 ) ■ (<Jj X a 2 ) = '\J(a 1 • a 1 ) ( o 2 • a 2 ) - (oj ■ a 2 ) 2
= \|all"22-a12=AJde<(aa/3)-
(1.7)
3
CHAPTER 1. GEOMETRY OF SURFACES
a = det(aa/3)
> 0 .
(1.8)
Let us next introduce the reciprocal metric coefficients by the expressions
a a/3 =
A ^
( y P ' 3 = cofactor of a Q/3 ) ;
j
(1.9)
then one easily verifies that the following relations hold
« a / 3 > = ^-
(L1°)
W e remark t h a t in general the vectors aa are neither unitary nor mutually orthogonal; when they are unitary and perpendicular to each other at a point, the coordinate system ua is said to be orthogonal at that point. It is often useful to consider, besides the coordinate basis ( a a ) , its reciprocal basis formed by the vectors
aa = a^a^
(afi = aa0aa).
(1.11)
Due to this proper definition and (1.10), these vectors satisfy the conditions:
aa-ap
= b%.
(1.12)
Moreover, if v is a vector tangent to ;f, we can write v = v"aa = vaa"
(1.13)
4
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where
( L14 )
•« = ««,»A
therefore, the contravariant components of v with respect to the reciprocal basis ( a a ) coincide with the covariant components of v with respect to the basis
(aa).
Starting from the quadratic form (1.5), which represents the metric on $ and is also called the first fundamental
form
of f, we can deduce all the
metric properties of the surface !f. Thus, if a curve 7 is given on if by the parametric equations
ua = ua(t),
te[a,b],
(1.15)
its length ^(7) is given by b
'(7) =
ds =
r*
dua du1*
*%■*%-"■
In the same way, let us consider a region a c $ obtained by varying the curvilinear coordinates over the set A = [a-pfc-J x [a 2 ,6 2 ] C 5t 2 . T h e area of the surface element da is defined by the relation
da = I du ax x du a2 \ = | ax x a 2 | dv}du2
which, owing to (1.7) (1.8), allows us to evaluate the area formula:
of a by the
5
CHAPTER 1. GEOMETRY OF SURFACES
•{adu^du1.
a
(1.17)
0
1.2.
T h e second fundamental form and the Gauss-Weingarten
equations.
T h e regularity hypothesis relative to $ assures (see (1.7), (1.8)) that, at least locally, we can define a unit vector field on If, which is orthogonal to If, by the relation a, X a ,
i n" == -ir - *
a
I l
W e say t h a t If is locally orientable field
2
x a
i-r i •.
(1-18)
2I
when we can associate to it this vector
which is orthogonal to f. Since the functions (1.2) are of class C ,
a • n = 0 a n d (1.3) holds, we can introduce the expressions
n n a af3= a3== ~ -aaaa-n,d ,P■n,0 = = aac,l3,f3a,0-n a
b
r
n = = bb b
= 0a P0a' a :.
a/3 3, 7 == rr«7/3 9 7 == aaa» -' aa/3, « 7 / 3 '' pa _ n aAp « aA r lA/37 . 1r£ 7 = a 0-y ~ \3-C
((1.19) !-l9)
2 ( L(1.20) °)
(1.21) (i.2i)
6
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
T h e quadratic form bag duadu" whereas Taa
is called the second fundamental
a n d Vg represent the Chrisioffel
symbols
form
of t
of the first and
second kinds, respectively. It is now possible to prove the Gauss-
Weingarien
equations ^,a = ^aa
7 +
6a/3n,
(1.22)
a-23)
»«=-*z«v where r ^ = 5 « 7 A K / 3 , a + «aA,/3-«a/3,A)-
(1-24)
In fact, since (aa, n) is a basis of the three-dimensional space 6 3 , we can write
" , a = ca °/3 + dc
n
■
(1.23) follows a t once from this relation, by noting that an ■ n = 0 a n d n n = 1 implies
n-na
= 0 . Similarly, by putting a/3) a = e^aa^ + f
0an
and recalling (1.19), we have fpa = bpa whereas a multiplication with a Q supplies t h e relation a^ a ■ ax = ayXe^a
which, owing to (1.20), (1.21), leads
(1.22). In order to prove (1.24), we take the derivative of the relations aaQ = aa- aa with respect to u and employ (1.20) to obtain
T
Pa\ + Ta0\ = aal3,\-
Cyclic permutations of the indices lead to the equations
(1-25)
7
CHAPTER 1. GEOMETRY OF SURFACES
l
Xfia + ipka
r
-
r
a
fiX,a >
a
aA/3 + Aa/3 = Xa,fi •
If we add to (1.25) the first of these relations, subtract the third one and take into account the symmetry properties of T a / 3 , we obtain (1.24). T h e great interest of (1.22), (1.23) resides in the following considerations. They represent an overdetermined system of five vectorial partial differential equations in two vectorial unknowns r(ua)
, n(ua).
One can suppose that,
under suitable integrability conditions, the aforesaid surface $ provided are known.
that the coefficients
aan and bn
system
determines
of the fundamental
the forms
T h e integrability conditions are relative to the right hand sides
and therefore they restrict the form of the functions aaa and ban. In order to determine
these
conditions,
integrability, two C
we
observe a
functions r(u )
that, a
, n(u )
owing
to
the
supposed
exist which satisfy
(1.22),
(1.23). From the obvious necessary conditions
" a , fiX =
r
, afiX =
a
a , A/3 >
(L26)
we deduce the necessary integrability conditions which have to be satisfied by the right hand sides of (1.22), (1.23). In fact, by evaluating the partial derivatives of (1.22) and taking into account (1.23), we have
«a,fix = (rZfi^x + Kfi,x-Kfii>x)% + (rZfib^ + ba0:X)n so t h a t (1.26) takes the form
8-
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
WcfiX ~ (*«£ bX ~ Kx tyK + (K0;X - Kx;P) » = « .
(1.27)
/here a
afSX -
v
aX, d ~
l
l
a0, X +
l 10' " 0X
1
aX
l
ia
Riemann ■ Chrisioffel tensor and
are the components of the
b
- b*0, X' -?Zx 6 7 / 3 "
a0;X
-rfc
b
oc-t •
The condition (1.27) is equivalent to the Codazzi-Mainardi
b
(3;X= aa0;X~
and the Gauss
equations
(1.28) (L28)
b
aX\0 a\\p -'
equationsC (1.29) (1.29)
b aXby R\ X=ba0=^-b a0 bX ~ baX b0 ■ R 0 "a0X
It is possible to prove that (1.28), (1.29) are not only necessary but also sufficient conditions for the integrability of the system (1.22), (1.23). In other words, the following theorem holds: class (y
are given which satisfy
one
only
and
one surface
quadratic fundamental
Remark.-
if the functions
admits
) , b a(u ) of
(1.28), (1.29), then there
conditions
which
aag(u
aag
forms to within a rigid
duadur,
b a duadu^
exists as
displacement.
When we bear in mind the symmetries of ban and T » , (1.29)
reduces only to two meaningful equations: b
a1;2 a\;2
=
-
b
a2\V a2;\-
CHAPTER 1. GEOMETRY OF SURFACES
9
Moreover, from the symmetry properties of the 16 components of
Rvaa~
given below
Rva0f— ~ Rav0-/
i
R
i/a0-/=
~ Rva-tP > Ruafi-/=Rf)-/i/a
(1-30)
it follows that 4 components do not vanish and only one is independent:
^1212
1.3.
=
-^2121
=
_
"^1221
=
^2112
=
"22"ll
_
"12
=
*"
(l.ol)
Normal curvature.
Other remarkable meanings are associated with the second fundamental form. Let 7 be a curve on the surface $ and r = r(ua(s)) be the equation of 7 (s is the arc length on 7). If x
ls
the curvature of 7 and fi its principal
normal unit vector, the curvature vector Jfc is expressed by the Frenet formula:
k = X» = fs,
(1-32)
where A = A a a a is the unit vector tangent to 7. By (1.22), the previous relation becomes
*=(^+W^H+M avJ »-
(1.33)
10
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let us define the normal curvature of 7 at (ua) e 7 as the quantity (see fig.l)
XnW
= k-n=bafi\aX<3.
(1.34)
fig.l This formula shows that all the surface curves passing through a point r(ua) have the same normal curvature at r(ua). If 7, ■, is a curve whose osculating plane is determined by A and n, we have
X„(A) =
k-n=x»-n=X
CHAPTER 1. GEOMETRY OF SURFACES
11
and therefore the normal curvature of 7, » coincides with the curvature. Moreover, from the obvious inequality
IX(„)| = | X M - n | < I x l we derive t h a t : among all the curves with the same unit tangent vector X, the curve j , \ has minimal 7 / 1 coincides curvature
curvature;
with the projections
moreover,
the centre of curvature
cn of
on its osculating plane of the centres
of
c of all the tangent curves to A .
1.4. E l l i p t i c , h y p e r b o l i c a n d p a r a b o l i c p o i n t s . M e a n a n d G a u s s i a n curvatures.
Let f be a regular surface of class C
and E2(r)
its tangent plane at the
point r. T h e first fundamental form, which is positive definite, allows us to regard E2(r)
as a Euclidean two-dimensional vector space. A basis for it is
given by (ava2) a
ad
a
^ **• Let
and the scalar product is defined by the metric coefficients
us now
consider the tensor baj3
This is a symmetric tensor over E2(r)
in the second quadratic form.
and consequently two real eigenvalues
if, and f 2i which may also be equal, exist along with an orthonormal basis formed by two eigenvector «j and v2 belonging to the aforesaid eigenvalues. These are solutions of the homogeneous linear system:
12
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
(6a/3 - V aap) v0 = 0 ,
(1.35)
whereas the eigenvalues satisfy the characteristic equation
det
(&a/3 -
which can be put into the form
(1.36)
where we have introduced the notations
H=\K
, K=\det{bafi)=\.
(1.37)
Let us refer the space E2{r) to the basis «j , v2 of the unit eigenvectors of bag . With respect to such a basis, bag is diagonal so that the quadratic form (1.34) can be written as Xtn\(ty
=
Vi(^ ) + ^ j ( ^ ) - ^ X( n )W does no*'
vanish identically (i.e. if both tp-^ and y>2 do not vanish), three cases may be distinguished: a
) fv
f2 7^ 0 have the same sign. Then the quadratic form Xf \(ty —
y>1(A1)2 + y?2(^ )
nas a
definite sign at r and consequently the normal
curvature always preserves the same sign on varying A. All the points of !f lie on the same side of the plane E2{r). In this case r is called an elliptic point. b) ip1 , ¥2^® have opposite sign. This means that the curvature along the two orthogonal lines, defined by Vj , t>2 have opposite signs. Therefore,
13
CHAPTER 1. GEOMETRY OF SURFACES
the tangent plane £' 2 (r) intersects f in a neighbourhood of r which is called a hyperbolic
point.
Moreover, two directions \i),^to)
exist along which the
curvature vanishes. These lines, which are called asymptotic
lines,
divide
£' 2 (r) into four regions; in each of which the curvature is in turn positive or negative. c) cp^ ^ 0, f
= 2
0- In this case,
^ , ,(A)= ^ ( A 1 ) . Therefore the normal
curvature has the same sign along any direction different
from
vanishes when A is parallel to A2. Then we say that r is a parabolic T h e graphical representation of these cases is given in fig. 2.
fig.2
A2 and point.
14
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
The eigenvalues of ba/3, which we denote henceforth by Xny the principal principal
curvature
directions
X(2)>
are
ca
Ued
of f at r whereas the eigenvectors of bag are the
of $ at r. From Cartesio's rule we have the relations
below X(l) + X(2) =
2
#
. X(l)X ( 2 ) =
By definition, the scalar H is called mean Gaussian
curvature.
A
(L38>
'-
curvature
of if and K the
The relations (1.37) 2 and (1.31) lead to the
Gauss
egregium theorem which is expressed by the formula
A' = ^ P , which proves that : the Gaussian
curvature
(1.39) depends on the first
fundamental
alone. This theorem has many fundamental applications which we can
form
not analyze for the sake of brevity. We just mention that all the intrinsic geometry of surfaces and, more generally, of Riemannian manifolds is related to it.
1.5.
L i n e s of c u r v a t u r e .
A curve j on f is said to be a line of curvature if its tangent vector at all the points is a principal direction. In particular, if X m = X(2)
at
eac
^ Pomti
CHAPTER 1. GEOMETRY OF SURFACES
15
then all the coordinate curves are lines of curvature (planes and spheres). It is possible to prove t h a t in a neighbourhood of any point r of a surface of class C
(k > 2) a local coordinate system exists such that the
coordinate
curves at
r
corresponding
are principal lines. It is also evident
that
the
coordinate curves are principal lines if and only if
a
(1.40)
l 2 = 612 = °
In such a system the following relations hold
ds
,a/3
u
0
ll
0
= Ojifdu ) + a22(du
JL
3
6
bn
0
0
6 22
a/3 =
22
)
a
u
ll
0
a
^ 22
6n h
n U - I (m. X. -21\ V - U 22 v -111 v -- :22 a 2 Vall a227 ' ll a22 ' X l ~ all ' X 2 - a 22
w ^=-i«^=-S^(%^) &
(log, WiTn).
( t t ^ / ? , no s u m m a t i o n over a and /?) r
«/3 = 2 ^ ~ % , i 3 = ^ 3 ( / o < ^ w ) ( n o summation summation over over aa)) .
16
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
1.6. Surface gradient operator and Gauss theorem.
Let u = uaaa + u n be a vector field on the surface if , not necessarily tangent to if. It is easy to verify that (1.22), (1.23) imply that its derivative is given by the formula below
« a = («7a " 6 > 3 ) «7 + ("fa + K^)
» -
(1-41)
where
«7« = «7« + I > A
(1-42)
are the components of the covariant derivative on if of the tangent field us = uaaa which is obtained at any point r e if projecting w on the tangent plane to if at that point. In particular, if « is tangent to if (u = 0), (1.41) reduces to the following formula:
u
,a =
U
(1.43)
^ a 7 + 6a/3^n-
Similarly, let T = TaPaa ®a0 + Ta3aa ®n + T3an®aa
+ T33n® n ,
be a double tensor defined on if. Then (1.22), (1.23) allow us to derive the relation 3 3 ^T ^« =f Cr£j - T_ T^/ 3 3- 6 7T -T ^^6f) ) ^a / 3®®a ^7
(1.44)
1
3
3 3 33
+ (Tf3 ++ T ^ 6 7 a -T^b -T ja-T6^)a^®n bP)a0®n + (Tf 33 33 3 3 / 3»3 f e6 + T 36 T 3 3 ++T T T 3^/ 6 &£) + {Tfa + T * 6-crv 6g) r»® n ® a~ a, + ((T 7Q - T a/3 Q / 3i )„ W® » ,
CHAPTER 1. GEOMETRY OF SURFACES
17
where y/3*y _ -
T ^ 7 .i-r^T^
+r j ^ , (1.45)
/33 rnn{53
T->/33 ,■ p/3 p/3 y 7 3 _ -7-/33
,a ~
,a '
0a-y 7
In the sequel we are interested in the case of a tensor T satisfying the condition n-T = 0,
(T3o = T33 = 0),
(1.46)
so t h a t (1.44) reduces to
. T / 3 3 f c 7 ) «« J J a T , a„, = = l(T^2 ; a -- J ° a J a / 3®» a„ 7 37 3 T f ++ T ^ 6 7 6a 7Q n ®na
3 +^ T
(1.47)
Let us now define the surface gradient of T as follows
VeT = a Q ® T Q ,
(1.48) .48) (1
and divergence md the surface surface divt •rgence
Vs-T
= a«.T,a
.
(1.49) .49) (1
Under Inder the condition n-T n - T == 00 , we get from (1.41) and (1.47):
1 3 V f «u 3 , , U 7 7 --22t # v„a -• uU =---v>7
(1.50) (1 .50)
18
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Vs-T=(T?2-Ta3b2)a1
+
(T?* + T^b^)n.
(1.51)
In order to point out the meaning of the previous definitions, we consider first a rectilinear system of coordinates (V) in &3 and the associated frame of reference ( O , ^ ) . Then « a = (u'ei),a = u',aei = u\j XU Ci =
\ j ( e j • aa)
U
and we conclude that « aa
e
i =
a
c ■ («* jej
® e t) = aa ■
V
"
(1 - 5 2 )
represents the spatial gradient of « along the vector
which is tangent to If. Moreover, for any vector «, the following
decomposition holds « = « s + n (n-u) ,
(1.53)
where « s belongs to the tangent plane E2(r) to f at r e if. Let us introduce the identity tensor / of S 3 and the projector tensor P on the tangent plane £"2(r) defined by the condition Pu
= us.
Obviously P reduces to the identity / on E2(r). From (1.53) we derive (1.54)
P-u = (I-n®n)-u. Similarly, for any tensor field T on f we introduce the decomposition T=Ts + n®{n-T) ,
n-Ts = Q.
(1.55)
CHAPTER 1. GEOMETRY OF SURFACES
19
By adopting the notation (a®b)-T
,
(1.56)
,
(1.57)
= a®(b-T)
(1.55) can be written as P-T=(I-n®n)-T
where P -T = T
if n - T = 0. On the other hand, it is quite obvious that
coordinate independent representations of / and P are given by
P = aa®aa,
I = e'
et ,
(1.58)
and by (1.54), (1.57) we a t t a i n the following relation
aa®aa
= e'®ei-n®n
(1.59)
.
The expressions (1.58) of P and / as well as (1.56), (1-52) allow us to write that
F • (Vu) = (aa ® aa) ■Vu = aa®(aa-Vu)
Hence, we can conclude that V s « represents We
want
now to prove that
= aa®u^a = Vsu .
the projection
the definition
ofVu
(1.49)
(1.60)
on E2{r).
leads us to a
generalization of the Gauss theorem to the case of surface vector or tensor fields which are not necessarily tangent to If. In fact, if w is a vector field on $, we have u? a
v-udl
,
(1.61)
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
20
where v is a unit vector which is tangent to f and orthogonal to the boundary df. By (1.50) we obtain [V s • u + 2H « - n) da - v - u dl
(1.62)
9$
Let T be a second order tensor field which is defined on f, satisfying the condition n-T = 0. By applying the result (1.62) to the vector field « =
T-v,
where v is an arbitrary constant vector , we obtain
[Vs-(T-v)
+ 2H n-T-v]
da =
vT-vdl.
Taking into account the condition n • T = 0 and the arbitrariness of v , we obtain the following formula:
V-Tdcr
, ( n - T = 0)
u-Tdl
=
(1.63)
df
For reasons that we make clearer in the sequel we need to extend the formulas (1.62), (1.63) to the case in which the surface $ is not regular along a curve T and the vector or tensor fields undergo jumps across T. We suppose that !f is divided by T into two parts $~, If"1" (see fig.3). By applying (1.63) separately to each of the two parts if+ and f ~ and adding the results we find that:
[V s ■ u + 2H n -«] da
v -u dl +
as
{vu}dl,
(1.64)
CHAPTER 1. GEOMETRY OF SURFACES
21
where
fig-3 {y •«} = 1/
■«
+ v
■u
and the meaning of other notations is self explanatory. If we denote by r = n x v the unit vector which is tangent to T and which determines an orientation on I\ we have v — r x n and then
i/"•" •«"•"+i/
■« —T^ -nT
x«T + r
•n
xu
22
But T+
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
= - T
= - T and we can then write:
v+
• u + +1/
provided that the sense of r
•«
=-r-[nx«]
is the same as the second term in the j u m p .
Finally, we can put (1.64) in the form below
[ V s - u + 2 # n - « ] da
v - u dl
(1.65)
T-lnxu}dl
dif Similarly, instead of (1.63), we have
V -T da =
v-T
dl +
{v-T}
(1.66)
dl
d$
or equivalently,
V-T
da
v-Tdl8$
T-lnxTjdl
.
(1.67)
Chapter 2
KINEMATICS OF SURFACES
2.1 Velocity of a moving surface. We call moving surfaces
embodied
surface in
the
a one parameter family of orientable regular Euclidean
three-dimensional
space
S3.
More
precisely, a moving surface $(t) is the locus of the points belonging to S 3 given by the equation
r=r(U\U2,t)
where the function
([/\t/2)eftc»2
,
r e C (fi x 3?) defines for every t g 5ft a regular
(2.1)
and
orientable surface of § 3 . In particular, these hypotheses imply that (see section 1.1): i) at every point of f(t),
the matrix {drljdU
) has rank equal to 2 so
t h a t the vectors S
A = ^
= f,A'
(
A = 1
.2).
(2-2)
which are tangent to the coordinate curves, are linearly independent, t h a t is,
23
24
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
they satisfy the following relation SjXo^O;
(2.3)
ii) the unit vector «=
"lX!2l ,
(2.4)
is normal to 3(t) and defines a continuous vector field on f(t).
All
geometrical objects relative to a moving surface will be functions of time. The metric coefficients a r A are given by a
rA — a r" aA ' whereas the reciprocal vectors are
(2.5)
a r = a r A aA ,
(2.6)
( 2
rA
) being the inverse matrix of ( 2p A ).
We remark that the parameters we have chosen in (2.1) are quite arbitrary and nothing prevents us from adopting at any instant t new parameters {u\u
) e Clt c 9? such that ua = ua(U\U2,t)
,
where the functions (2.7) are invertible for any t. If the moving surface
(2.7)
f(t)
represents the actual configuration of a bidimensional continuous system S, that is when it is material, we can preferably adopt (U ) as parameters to describe if(<). Then these parameters are called the material or lagrangian coordinates U
which correspond to a coordinate system in a reference
configuration $(t0) of $(t). Differently, when if(i) is a geometrical surface,
25
CHAPTER 2. KINEMATICS OF SURFACES
which is not connected with moving particles, the parameters U1,!/2 are quite equivalent to u\u
[ 1, 2, 3, 4, 5 ] .
Let ?(t) be a moving surface and (U ,U ) , (u , u ) two arbitrary parametrizations of lf(£) related to each other by a transformation (2.7). Then we have r=r{u1,u2,t)
= f(U1,U2,t)
.
(2.8)
It is possible to associate to the points of $(t) a velocity depending on the parametrization by the definitions
'=($„.
d
*
~\dt)uA-
(2.9)
From (2.8) and (2.9) we see that
C =
or
( W ^
\di)u<*
+ r a
'
VdTJjjA '
+ ^-—m v *" ($V*-
2 io
<- > (2.10)
If we apply the decomposition (1.53) to c and c, the relation (2.10) can be written as c„, +
(c.n)n=cs
+
(c.n)n
+
(^£)^« '[/ "
so that we finally obtain d
s
-A%-)u.'"
c = c - n = c - n = c„ .
(2.11)
26
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
The second of these equalities shows that the normal component of velocity of a moving surface is independent of parametrization.
On the other hand,
(2.11)j, which can also be written in the form ?A C
s
ML
a
duA
"a -
-c«a C
,§vZa
a
s
a
a + Qf
a >
shows that the surface velocity becomes purely normal, i.e. c = cnn , with respect to a parametrization ua(U , t) which solves the following partial differential equations dua
dt
8ua
_
- A /rrl C
~5t/A
*
[U
rr2x
'U
>■
For this reason this type of parametrization is called a normal
parametri
zation for ^(<). If the surface if (i) is given by the equation f(x, t) = 0, where x = r(ua, t), we have
\dt)x
+ v/-|'6r\
KdtJua
and thus
<%b|
Vf\n-c--= 0 ,
3/
c
"=-p§5r-
(2-12)
Now we want to evaluate the time derivative of some kinematical quantities for a future use. We first have a
and this in turn results in
«
= f
!a
— c ,a '
27
CHAPTER 2. KINEMATICS OF SURFACES
,i
a/3 = c , a - a / 3 + a a - c , / 3 ■
When we take into account (1.41) , the previous relation assumes the form:
^ =2 ( { { ^ ) - W s 2 ^ '
(2-13)
We are now in condition to evaluate the time derivative of a = det(aaa). In fact, it is evident that da a0 ■ ■ a— ®a aa(} = aa a8 ^ aa0 . _ (9(7.
r
Thus recalling (2.13) and (1.50) we obtain a = 2a(c? a=2a(c? 2aVss.c■ c ==22arf < < .a. a-2HcJ a - 2Hcn) = 2aV
(2.14) (2.14)
From this formula we can now calculate the time derivative of the area element da and of the normal unit vector n to 'S(t). By (1.17) and (2.14) we have jfida)
= 7 ^ duldu2 = ^aVs-c
du^du2 ;
or jL(da) = Vs-cda
= 2r]aa.
(2.15)
Moreover, the time derivative of the orthogonality condition n-aa = 0, implies that n • aa = -n-c
a
c
and by (1.41) we are then led to the equation
n, a
= -n-aa-6^C/3.
28
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
But nn-n= ■n
=
a ,41)): 1 => = > n -•n n = 0=>-hh = (h-a-a)« ) aa so that we obtain (see (see (1 (1.41)): = 0=^ =
( * ■
= -- (( cc „n ,j 0„ ++^&%)«" ) a « = = ~ (»(»•«=,«) » . c J a «° °.»« =
2.2.
(2.16)
T h e Thomas derivative.
Let F — F(r,t)
be a time dependent field (scalar, vector, etc.) defined on
f (<). If r = r ( u a , t)
is the equation
of the surface,
we introduce
the
representation a F = F(r(u F(r(ua,t),t) ,t),t)
( ( = '§(u '§(u *,t), *,t),
(2.17)
where the function ?F depends on the chosen parametrisation of f(t). definition, the Thomas
derivative
or the displacement
derivative
By
of F is the
following limit
6F — Urn F ( r + c n n At,t + At)- -F(r, t) At fa At-»0
(2.18)
In other words, the T h o m a s derivative denotes the rate of the change of the field F with respect to an observer moving along the normal to $(t) with the velocity c n . W e note that the limit (2.18) does not depend on the choice of parameters ua owing to the independence of c n of this choice. It is evident t h a t (2.18) implies that
CHAPTER 2. KINEMATICS OF SURFACES
29
f = (f)r +C""-Vf-
<2-19)
On the other hand, it follows from (2.17) that
F -
<%.+<
■VF-- =
(dt)u«>
and consequently (2.19) becomes
f = (ft-<W*.
(2.20)
But (1.60) and (1.57) imply that
V , 5 = V F - n ® (n■ VF) => cg • V s 5 = c s - V f .
Finally, we may obtain a new expression for (2.20) in terms of parameters
TH^a-V^^'-V^ 5 -
(2-21)
In view of (1.49) this last relation can also be written as
f = (f)u«-<*.-
<2"22)
30
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
2.3 The velocity of a moving curve on a moving surface. Let T(t) be a moving curve which remain always in $(t). If $(t) has the parametric representation r = f(U , <), the parametric equation of
T(t)
becomes r = r(UA(ti,t),t)
=
ip(fi,t).
The velocity of T(t) in the parametrization we have chosen is given by
c=
t a r i „ = \§t.)VA + 5^ vw)^
so that in view of (2.9) and (2.2) it can be put into the form
c ==«♦(#)„ «A •
(2.23)
If we introduce different parametrizations for f(t) and T(t), that is if we use the new parameters ua = ua(V
,t) for $(t) and /J, = p(A,t) for T(t), we have
a different equation for T(t): r = r(ua(UA(fi(X, t), t), 0) - f(UA(fi, t), t) , and the velocity C of T(t) becomes now c-(dr\
(dr\ _(dr\
\dtJ =\dt) aa °<%\ x
uu
dr j (du<*\ ,_d^f(dv^
+ +
aa
a
+
dua ((dUA,du ) A
dt
(dUA\(d»\ \ ] ((dV±\.(dU±\(dtA\\
du \1\ \ dt \dp )\dt)J du dt 7y^A at/A\S^ dtb)»]\d}i) \dt)x) ( r [/A 0U
=-(¥u°«4ia+^(w)>-
31
C H A P T E R 2. KINEMATICS OF SURFACES
W h e n we note that ( dU
/dfi) a^ is the tangent vector f to the curve T(t) in
the parametric representation U
= U (n,t),
we obtain
A Ss A ++J°£\9.f . a JW \ a. c ==-(¥h°° dt (1)©/" ♦ ( # ) )„ . A
C =:eJ*£\
(2.24)
By comparing (2 .24) and (2.23) and taking into account (2. 11), we deduce that C-C = © r .
(2.25)
prof ection of This result resu It shows that the projection ofCC along the unit vector v tangent
m 'Ht)
and
and orthogonal
f(t) m
to
to T(t) T(<) is independent indepi indent of bi both ylh parametriza parametrizations tions of T(t) r(t)
Csauu
= C = cssu.- v
.
(2.26) (2.26)
Moreover, from (2.23), (2.25) we have Cn = C'n = cn in agreement with the condition T(t) c S(<).
2.4 Time derivative of a volume integral.
In this section and in the subsequent sections of this chapter we will state some general formulas to evaluate the time derivatives of volume and surface integrals.
32
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let C be a continuous system and (2.27)
x = x(X,t)
its equation of motion, where X is any point in the reference configuration C+ and x is the corresponding actual spatial position in the configuration C(<) a t the instant t. We denote by V the velocity of the boundary dV(t) of any region V(t) enclosed in C(i), by v the velocity field of C(t) and by N the normal unit vector t o dV{t).
Our actual purpose is to evaluate the following
time derivative
dt V(t
(2.28)
f dv ,
where f(x, t) is any field ( scalar, vector, etc.) associated t o the continuous system C. Now, it is quite obvious that
A.
dt V(t)
f dv =
±(fdv)
v(t)
N-(V-v)f
+
da
(2.29)
dV(t)
W e note that the second integral on the right hand side represents the flux of / across dV(t) due to the combined effect of the^motion of C and dV(t).
On
the other hand it is well-known that
dt
(dv) =
(2.30)
V-vdv
and (2.29) can be written as follows
A dt V(t)
f dv =
[f + fV-v]dv V(t)
+ dV(t)
N- {V-v)f da
(2.31)
33
CHAPTER 2. KINEMATICS OF SURFACES
The following subcases of (2.31) are interesting: i) V(t) is material, that is VN = vN
; «) v(t)=-V
is fixed. In the former case we have: d
f dv = [f ++ [f dt *{'"•=, V(t) V(t) VV(t) V(t)
fV-v]dv. fV-v]dv.
(2.32)
V t)
whereas in the latter (2.31) becomes d dt
r
f dv == V
df dv dv Mdv --Tt
(2.32)'
V
In the sequel we need a generalization of (2.31) to the case in which / undergoes jumps across a surface £(£) moving with a normal velocity c . To +
attain this goal, we suppose that V{i) is divided into two parts V ~ ,V £(£), where y
+
by
is the part toward which the normal unit vector n to £(£) is
directed. By applying (2.31) to V _ a n d Vr + and adding the results, we have
[1]:
dv i\'V(t) =l V(t)
[f + + fV.v]dv+ fV-v]dv t) V V(t)
+
*
dV (t) dV(t)
f(VVN-vNV)daf ( N~ N)
IfU]IfUldcr, da ,
d(T
~
E(t) a *)
(2.33) (2.33)
where we have notations V V introduce 1 11 \j 1 KJK.I IJV_.(_ the U11V, 11 K.' UC* H U l l O = (V-v)-N U±=c=n-v± ((V VNj-v vN-)t . ^ ) = (K-i»)-JV , tf* « « - « „ * •. We want to emphasize that : i) in all the formulas we have deduced so far only the normal velocities Vjy and c n which do not depend on the parametrizations of dV(t) and E(t) appear ; ii) (2.33) is also valid in the more general case of a discontinuity surface dividing V(t) in many parts
34
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
provided that the boundary <9E(tf) lies on dV(t) (see fig. 4). To verify this it suffices to apply (2.31) to each of the aforesaid regions.
fig.4
2.5
T i m e derivative of a surface integral.
Let r = r{ua,t)
(2.34)
be the equation of a surface E(£) moving with velocity c and cr(t) a region of £(£) whose boundary da(t) is a curve which moves on E(<) with velocity W . In this section we want to evaluate the time derivative
35
CHAPTER 2. KINEMATICS OF SURFACES
F da
dt
(2.35)
where F = 5 ( u a , t) is a field ( scalar, vector, etc.) on E(<). Instead of (2.29) we now have
d dt
F da = = '
*{ t)
j-t{Fda) j-t{Fdo)^
v- {W-c)Fc)F dl .
(2.36)
da da(t) (t)
Owing to (2.15) and (1.50), this relation can be written as
d dt.
a(t)
F da =
-c -2Hc -2Hc IHcF] F]F] da + + f[($L+".-« s
ss
nn n
*(t)
- c)F dl . w{W-c)F v-{W-
fl_ 11\ dcr(t)
(2.37)
By recalling (2.21) we finally obtain
d F)-2Hc F da = [tg+Va.(ct®F)-2Hc nF]da+ nF)d
jf d
v{W--c)F v(W-
dl . (2.38)
It is very important to remark that the time derivative of the surface integral explicitly depends on the surface parametrization because of the presence of c s in both integrals on the right hand side of (2.38) . This is due to the fact t h a t the velocity cs is a priori physically meaningless. In other words, if we want to use (2.38) in physical balance equations, we are compelled to introduce a parametrization on the surface to which a physical velocity is associated. For instance, if T,(t) is material, i.e. if it is a twodimensional continuous system, we can use as parameters the Lagrangian coordinates of the particles of £(<) so that c e denotes their tangential velocity to E(<).
36
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Formula (2.38) can be extended to the case in which F undergoes jumps across a curve on E(<) which moves with velocity C and is singular for the surface E(£). We start with the hypothesis that £(£) is divided into two parts £~(£) and E + (<) by the curve T(t). If we apply (2.38) to each part and add the results, we obtain
[^
F da =
dt
Vs-{cs®F)-2HcnF]da
+
'(*)
+ j v(W-e)Fdl+
(2.39)
{u- {C-c)F}dl ,
r(t)n
a
where we have again used the notations of section 1.6. By the same procedure of that section, we can transform the last term of (2.39) as follows
{v {C-c)F} dl= r(«)n
r-lnx{C-c)F}dl
.
(2.40)
r(t)n
We want now to derive an interesting well-known consequence of (2.39). Let lf(f) be a material surface moving with a velocity field v and let F be a scalar field of the form u-N , where N is the normal unit vector to f(t). In this condition the integral u-N da , T{t) denotes the flux of the vector u across the surface 'S(t). From (2.21) , we find with W = c = v S
-£+Vs-(vs®F)-2HvNF
CHAPTER 2. KINEMATICS OF SURFACES
= F-v-v -F+ F(V F(Vss-v -vas-2Hv -2HvNN) ) s-V sFaF s-V
37
+ vvss.V -Vs3F.
If we take into account (1.50), we can immediately write the following equality 6
-£ +
Vs-(v3®F)-2HvNF =
F+FVsv.
But (1.41) and (1.52) imply that a a Vssv ■ v==aaav, V aa ■aVv ■=aatr[{a = tr[(a aa-Vva ® aaj ®■ajVv]■ Vw] a = Q■ v, =
and (1.59) allows to write W + + VVs-(v -F2HvNF ==FF++F(V-v-N-Vv-N) F (V • v - N ■ Vv ■ N) f£ ®F)-2Hv . .(2.41) (2.41) s • s(v s ® F) N .52)) (1.52)) If F = «• u ■NN we have (see (2.16), (1 d -^-(u-N) == u-N uii ■■NN + «u-N=ii-N-(N-v -(N■N) = ■N) v■TV N = = iiu ■N ■N- -(Na) dt (« = «u-N-(a = ■N-(aa--Vv-N) V v N) Uua
V y
a)
u ua aaa
and (2.41) becomes f,V
_
^+V &F + v s • (v, ®F)s-(v,®F)-2Hv NF
VvN++u-N 2HvNF ==ii-N-u ii-N-usu-N-u N -VvN ■-VvN Vv ■ N + u -u-NV-v TV V -Vw■ v sVvN-s-VvN-- uuNNN _= ( (^du + + vv-Vu-u-Vv V u --« • Vv + uV-uV-v)-N . ) . TV ~\ dt = ( | | +V Vx(uxt>) x ( u x t > ) + wV-u)-JV. wV-u)-TV.
(2.42)
38
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
On the other hand, the jump in (2.40) can be written as
-T-INX(C-V) -T-[NX{C-V)U-N]
u-N}= = -T-[UNNX(C-V)] -T-\UNNX(G-V)1
= -T-[{u-ua-r-l(u-u )x{C-v)] s)x(C-v)} = where « s is the projection of « on the tangent plane to f(t). By representing C-v
in the basis (r,v,N)
and noting that CN — vN , it is easy to verify
that -T-[Nx(C-v)
«.JV1= - r - [ u x ( C - v ) ] .
(2.43)
Collecting the results given by (2.42) and (2.43) we finally obtain d u- N dcr = [ | ^ + V x (w x w) + v V •«] ■ N da dt\ J(«) J(t)
r-[ux{C-v)]dl.
-
(2.44)
r (0
2.6 Two time derivation formulas derived from (2.39) and (2.44).
The interest of (2.39) and (2.44) has to be found in their use in the general balance law which we consider in the next chapter. However, (2.39) (2.44) cannot be used in their present form. In effect, to use them we need to make explicit the velocity W of the curve da(t) in (2.39) and the velocity C of Tit) in (2.44).
- 39
C H A P T E R 2. KINEMATICS OF SURFACES
Let V(t) be a material volume contained in the actual configuration C(t) of a continuous system, TV be the normal unit vector to dV(t),
£(<) c C(Z) be
a singular moving surface and T(t) C £(<) be a singular curve. If we want to apply (2.39) to the surface cr(t)
=
V(t)nT,(t)
velocity W of dcr(t) to the velocities v of dV(t)
, we have to relate the and c of S ( i ) (see fig. 5). T o
begin with, we remark that the vectors TV , n and the vector i / s , which is orthogonal to dcr(t) within the tangent plane to a(t)
are coplanar at any
points of dcr(t). In fact, TV is orthogonal to all vectors tangent to dV(t)
and
to the unit vector tangent r to dcr(t). But both n and J / S are orthogonal to r and to each other so t h a t the coplanarity of TV , n , J/^ along dcr(t) is held.
fig-5 Let f(x,t)
= 0 and g(x, t) = 0 be the equations of dV(t)
respectively. T h e equation x = r(\,t)
and
S(t),
of the curve dcr(t) satisfies the system
40
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
f(r(X,t),t)-- = 0, ff(KM),<) == 0. By differentiating with respect to t we have
v/.^+§H
v , . ^ + | = o.
On the other hand, owing to (2.12),
df dt
V
N ~ -
dg dt |Vff| '
C
[v/l
n~
and the previous equations imply WN = vN
W
= c
(2.45)
When we choose the basis (T,I/^,N) (r >"£> f) ,, we have W=W W=WTT TT + W W' v-^ ++ WW u *2 nnnn and it is very simple to verify that (2.45) leads us to the formula
w= =W ">
T ++{l^N ( ^
W
N)
c
„»-■ *fc++C n» ~--cCnn ^W)■NJ "E
(2-46)
c n n • AT )
(2-47)
Consequently, we find that (W
e) • * £ - ( « #
In view of this expression we can write
1^
c-i/E.
(2.39) in the situation we are
considering as follows d dt
fFdada=
f [7f+V =\^+Vs-(c 1Hc F] da nJnF]d<7 s®F)- s-(cs®F)-2Fc
" • ( < )
+ da
f
(0
F
KVN
c n n-JV)
" ■ ( * )
v
\
N
c-vx]
dl
r r(t) no-
rr-lnx{C-c)F]dl .[nx(C-c)f](f!
(2.48)
C H A P T E R 2. KINEMATICS OF SURFACES
41
where a{t) — E(<) D V(t) , V(t) any material volume and T(t) is a singularity line on £(£). Now we want to apply (2.44) to the case in which the moving curve T(<) on the material surface if(i) is obtained by intersecting $(t) with a moving singularity surface E ( i ) ( see fig. 6). First of all we note that r = i/ E x n , and then we have -r.[tix(C-»)]=
-i/E-[nx(«x(C-i»))]
where C is again given by (2.47).
fig.6
Chapter 3
BALANCE LAWS FOR A CONTINUOUS SYSTEM WITH AN INTERFACE
3.1 General balance laws.
The weight of balance laws in Continuum mechanics is well-known. T h e mass conservation and the balance laws of momentum, angular m o m e n t u m and energy constitute the fundamental principles to derive the corresponding local equations as well as the boundary and j u m p conditions of the fields which describe the evolution of the system C(t) we are concerned with. In their integral forms the balance laws are postulated to be valid for every fixed or material region of t(t)
(see, for instance, [1]). More recently [6,7,8,9],
it has been suggested that the above mentioned balance equations can be postulated only for the whole system C(<) but not for any of its arbitrary parts (Nonlocal Mechanics). In this chapter we propose to use Continuum Mechanics to describe the evolution of a system C(t) with an interface
E ( t ) . In such a system £(<) does
not reduce to a singular surface for the volume fields which C(i) carries in its
42
43
CHAPTER 3. BALANCE LAWS
evolution. On the contrary, £(£) has to be regarded as a nonmaterial which
is
able,
however,
means t h a t
£(£)
can foresee
that
to transport
thermomechanical
surface
properties.
This
is itself a physical system in interaction with C(t) . One this model is appropriate to describe phase
transition
phenomena; in fact, in such kind of phenomena two or more bulk phases are separated by very narrow layers in which the volume fields vary continuously but sharply. This layers can be regarded as discontinuity surfaces for the volume fields provided that some suitable surface fields are associated with these surfaces in order to take into account the mass, energy etc. contents of the layers they represent. It is quite obvious that the layers do not contain the same particles of the system during the phase transition and then the representative surfaces are not material. W h e n we try to use the formalism of Continuum Mechanics to describe the evolution of the system with an interface (C(<),E(<)) we are faced with nontrivial difficulties. In fact, we need to make clear: i) how a nonmaterial surface is able to transport thermomechanical properties , it) how we must modify the classical balance laws when there is a convective flux of quantities associated with £(£) across the curve £ ( / ) Pi dV(t) material
volume V{t).
on the boundary of any
In this section, taking for granted that
we can
associate physical quantities to a nonmaterial surface £(<), we propose a general balance law for the system (C(i),£(2)). W e postulate t h a t any balance law for the system (C(<),£(<)) has the following general form:
44
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
dt\
f dv +\-
V(t)
F da) =
8V{t)
Rda+j
N• (p d(T +
r dv +
V(t)
i/ s • * dl
d
F[(vN-cnn-N)L-±7f-c-vE]dI,
(3.1)
dcr(t)
where V(t) is any material volume,
fig-7 The following remarks about (3.1) are worthwhile:
due to the
45
C H A P T E R 3. BALANCE LAWS
i) the last term in (3.1) has never been considered in literature ( see, for instance, [3, 4, 5]) up to 1986. In paper [10] it is considered only in the case of a fixed material volume V(t) (i.e. vN — 0) and independently in [11] in the complete form; ii) if the interface is material, for instance in the case of two bulk phases of immiscible fluids, the last term in (3.1) vanishes since W = c = v (see (2.67)); Hi) if V(t) coincides with the whole volume C(i) of a system with an interface, the balance law (3.1) assumes the classical form:
^ ( j fdv+l C(t)
Fda) = | N-tpd
aC(t) • $ dl +
J/ E
9E(t)
r dv
C(t) i i da .
(3.2)
E(t
In fact, <9£(<) moves on 9C(i) so that again W = c on 3£(£)
and the
convective flux of F across 3E(<) vanishes. We will come back to this remark when we propose to employ the Nonlocal Mechanics to describe phase transitions in crystals. It is now possible to derive the partial differential equations and jump conditions which are equivalent to the integral law (3.1). To this purpose we begin with recalling the classical Gauss theorem for a field
N-
\V-
n-{ip}da. "
(3.3)
46
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
When we take into account (2.33), (2.48), (1.67) and (3.3) we obtain
[/ + / V - w - V y j - r ] dv V[t) 8
+ | [ -£+Vs-{cs®F)-2HcnF-Vs-*-R-UU
+
n-tp\]d
T-lnx({C-c)F
+ <S>)} dl = 0
lit) where y(t) = T(t) fl cr(t) and T(t) is the moving line of discontinuity on £(£) (see fig. 7). T h e arbitrariness of V(t) and the regularity of the fields in C(t) - (£(£) U T(t))
lead
us to t h e following
local equations
and
jump
conditions: / + / V - v - V y > - r = 0,
^ f + V s • (c3 ® F) - 2HcnF
mC(t)-E(t)
- V s • * - R - [fU + n.
T-[nx((C-c)F + 4)] = 0,onr(t).
(3.4)
Our axiom that all the balance laws of a continuum with a n interface have the form (3.1) is valid only in the absence of electromagnetic
fields.
When these fields are present, we have to take into account balance laws of the form (seefig.8)
j-A 1(t)
u-Nda=
| a-r dl + af(t)
b- N da + \
i(t)
(3.5)
r,(t)
where $(t) C C(<) is a n y material surface, rj(t) = E(<) n f(t) a n d r t h e
47
C H A P T E R 3. BALANCE LAWS
tangent unit vector to df{t).
Moreover, « and a are vector fields in the
volume C(t), 6 and ib vector fields defined on £(*). By noting that r =
"X
xn
and applying (2.44) and the generalized Stokes theorem we obtain
a if (t)
[a]•r dl ,
V xa- N da +
a-T dl =
f(t)
vU
we have
fig-8
and
m
(du + \7 x (u x v) + v Vu - V x a - 6) • N da
(nx{ux(W-v) Ldl=0, ( n x | u x ( W --v) + + aa] | ++*J • k)'V f £ dl =U,
-
M n(t)t)
(3.6)
48
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where the velocity of r){t) is given by (2.46). The arbitrariness of the material surface f(£) leads us to the following local system | ^ + V x ( « x « ) + v V « - V x a - 6 = 0 , inC(<)-£(t) ,
(nxlux(W-v)
+ a} + k)-vE
=0,
on £(<).
(3.7)
This last condition is satisfied for every v^ tangent to £(t), however, it does not imply that the tangent vector which multiplies J/ S will vanish since W depends on i/ s ( see (2.46)).
3.2. Mechanical modelling of nonmaterial interfaces.
In this section we propose to make clear the model we will take into account in the sequel as well as to supply a possible interpretation of the mechanical properties of a nonmaterial surface. We have already said that a continuous system with an interface consists of two continua which occupy two adjacent regions Cj and C2 separated by an interface £ ( for the sake of simplicity we do not explicitly write
the dependence on time t). £ is
regarded in turn as a two-dimensional continuum with assigned proper constitutive equations and it is presumed either to be permeable or not and
49
CHAPTER 3. BALANCE LAWS
possibly adsorbing.
When E is permeable, we assume that the matter crossing
E, for instance from Cj to C 2 , is to be described by the same constitutive equations of C 2 . If E is not permeable, the model describes a system made by two solids or immiscible fluids occupying two adjacent regions. The task of the two-dimensional system E which represents the contact surface of the aforesaid regions is to take into account possible interfacial phenomena. In particular,
E could be a real two-dimensional continua (oil)
interposed
between three-dimensional continua (water and air). When E is permeable, the hypothesis t h a t m a t t e r changes its original constitutive equations into those of the continuum occupying the arrival region makes the model able to describe the state change. In order to complete the model (Cj,C 2 ,E) we must define its balance equations and the set of constitutive equations for each of its component. The former
are
momentum,
represented energy
by
mass conservation
balance
laws
as
well
as
and by
momentum,
angular
the
law
second
T h e r m o d y n a m i c s . In other words, we need to define the quantities
of f,F,...
which appear in (3.1), (3.4) for each of these laws. The most natural way to attain
this
objective
consists
in defining
the
volume quantities
as
in
C o n t i n u u m Mechanics and in associating corresponding quantities to E ( den sity, velocity, stress,...). However, two objections can be raised to this widely followed approach [2,3,4,5]. First of all, when E is not material, the meaning to a t t r i b u t e to the velocity of the particle which instantaneously lies on E is ambiguous. In the second place, the local angular m o m e n t u m equation which is derived in this way is not Galilean invariant. In order to overcome the above objections we want now to propose a model which leads in an almost
50
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
natural way to the proper quantities to introduce into the balance equations . We recall that the interface between two contiguous systems is really a very narrow layer formed by matter having different characteristics from the adjacent continua. This suggests to regard (
in details, we substitute the
layer with a surface E C AKy and then we associate surface fields to E which are averages of corresponding volume fields. This procedure lays itself open to a fundamental criticism. In fact, when we substitute the layer with a surface we extend, at the same time, the bulk phases to fill the layer which is now e m p t y . But this operation adds an amount of mass to the system which is of the same order of the mass we have concentrated on E. In other words, the system (C-[,C 2 ,E) is not equivalent to Cy. For this reason Gibbs [12,13] suggested to localize E at a suitable position in the layer and to put on it a surface mass distribution (possibly negative) by requiring that the total mass remains fixed. However, in dynamical situations we would localize E in such a way as to conserve also, besides the mass, the m o m e n t u m ,
angular
m o m e n t u m etc. and this proves to be impossible. To overcome this difficulty, we proceed as follows. i) Let C( be a family of not equivalent systems C(, where /-+0, including the system Cy we have started from, such that in each layer AV(
of C, the
content of mass, momentum,... is equal to the corresponding one in AVy . In this way, / determines the order of magnitude of the mass associated
51
CHAPTER 3. BALANCE LAWS
with the interface and moreover, when /-»0, the difference we have between Cj and the system with an interface obtained by concentrating mass, momen t u m , etc. on a surface S in AV ( goes to zero.
Then, we postulate that the
balance equations of our system (Cj,C 2 ,E) are equal to those we derive for C ; , when /-»0. It is evident t h a t this procedure rests on the assumption that, in the first approximation, the detailed distributions of mass, m o m e n t u m , etc. in the layer are not significant. If cl C C( is any material volume, a general balance equation for the system C ( can be expressed as
d
f dv = cI
N -cp da + r dv , c/ 9 c,
(3.8)
with the usual meaning for the symbols (see section 3.1). Concerning the system Cj we suppose that (see fig. 9): ii) C, = C 1 (/) U C 2 (/) U AVt parallel surfaces !f-(/) , i= 1,2,
, where the volume AV
t
is bounded by two
moving with the same normal velocity cn
and by the surface fy, belonging to 5Cj. Moreover, the distance / between the surface !f,-(/) is independent of time and 0 ( is orthogonal to both ^j(/)'s along the boundaries. Let Cj be any material volume such that ct n AVj ^ 0 and verifying the same conditions listed in ii). We denote by small letters the corresponding p a r t s of C[. Then we can write the first and the third integral of (3.8) as the s u m of the integrals over the three regions Cj(/), c 2 (/) and Az^ and apply Fubini's theorem to the integrals performed over Av ( , where Av{
can be
52
-PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig-9 represented by
A„, = U J(r) re
Here a C £ and £ is a surface which is equidistant from S-(l) whose unitary normal vector is n, while J(r) = (r + £n, - 1/2 < £ < 1/2). Let us now introduce the notations Ft(r, t) = j fj(r, 0 dZ, where
r
*( )
J(r,£)dtr{r)=dtr{r
rj(r, £) dl ,
R,= H*)
+ Sn).
We remark that assumption ii) implies that the limits of F[ and R{ are finite
CHAPTER 3. BALANCE LAWS
53
when /-»0. With these notations we have d dv = ft fdv+it dt ' f c(l) / ) U c 2 2( (0 /) Ccl x( (/)Uc e(l)
Ft da , '
(3.9) r dv = c(l)
rrft) +
ify dcr .
c x (/)Uc 2 (0
If we put <9c'(/) — dc(l) - u>(I) and note that
u(l) = ( J I(r) ,
AT = i/E on w(/)
the flux term can be written as
N -( d/
N -(p dv flc'(r)
dm
(3.10)
a
where */(»■."£)
=
p(r + fn)J(f,r,i/ s ) d£
and dl = dl(r, v-^) is the line element on the curve da, i/ £ is the unit vector normal vector to da and tangent to E and finally J(£,r, i/j>)dl(r, i/ £ ) = dl(r+£n, i/ s ). We note that the
vector field
normal v-% when / ^ 0 so that the Gauss theorem is applicable to the second integral on the right hand side of (3.10) only in the limit /->0 (when /->0, J-*\ and $; is independent of i/g).
54
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
We now have to substitute (3.9), (3.10) into the formula (3.8) so that it now reads
ft]
fdv+^lr,***
C l (/)Uc 2 (7)
«
i/£ ■ % dl
N -
Qc (I)
3 <7
+
R{ da . R,d
rrdvdv +
(0U C el m ucc2 (m( )
(3.11)
"
It has to be remarked that in the above formula both cAl) and cdc' c' (/) are regions depending on the parameter /. To apply (2.33), (2.38) in (3.11)
we
must define the tangential velocity c gs of a. To this regard we assume tthhaatt Hi) each parallel surface to E contained in AVl c
has a tangential velocity
r
particles «s ( ' 0 which is equal to the tangential projection of velocity of the particles
lying on it at any instant. The condition that w ; is orthogonal to SAl)
and E £ implies t h a t the
all these these velocity projection Vss(r)— ( r ) = (w(r + ^n) • i/ s ) j / ^ is uniform along I(r) ; all Cs = cc^ = Vs and then we have considerations permit us to put in (2.38) C,
|
[f fV-v]dv[/ + / V-v] dv- |
cC
ll(( / ) U c 2 ( 0
J 2 c- vfx(cnJ [/ 2 (
- vln)Jt]
dcr da
Ss
«(0 i (')
). l)-HcnF [6-^+V -Hc + [-^+V s.(V s®Fll,®F nFt]da s-(V l]da
= =
c
N -
(I) dc dc'(l)
'
+ c x (0 IJc 2 (/)
r dv +
Rt da , r
(3.12)
55
CHAPTER 3. BALANCE LAWS
where fi
is the value of the field / on the surface 5 t ( / ) , c n is the normal
speed of S^l)
and E, Ji — dajda;
moreover, dai and da are the correspond
ing surface elements on 5,(/) and E respectively, vin=vi
■ n and v- is the value
of velocity field on 5,-(/). If we now calculate the limit of (3.12) when /->0 (we suppose that for all functions F j ,
/ + / V v - V - p - r = 0,
6
-£+Vs-(V (V, ® F)nF-V-*-R --2Hc nF- -V . $ . R s •s®F)-2Hc St + v -[/E/
+
n^]
+
{ h ™ § - f } = 0.
(3.13)
In (3.13) F and R denote the limits of Ft and Rt respectively. Apart from the last term within brackets, equation (3.13) 2 coincides with (3.4) 2 but now we have the identification of F , R with suitable averages along the normal n to E of the volume fields / and r. To continue further, we now assume that iv) there exists a value /* of / such that F * = Ft for any / < /*. This assumption means that the family C(/) of systems which we are considering in our limiting process not only has the same amount of mass, m o m e n t u m , etc. in the layer AV[ (see assumption i) but also in each small "cylinder" da(r)x iv)
cannot
[-1/2,1/2].
be validated
In effect, we will show that the assumption
for all the averages Ft
because they are
not
56
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
independent of each other. In other words, iv) has to be employed under the condition that it does not lead to contradictions. If we refer to the mass conservation, /
is the mass density p and
consequently ps = lvy,l
= | pJ di .
(3.14)
I{r) Similarly, for the surface m o m e n t u m
P-lim
P we have
P.-Urn
(3.15)
pvJ df I(r)
so that we can define the surface velocity as that of the centre of mass of the particles instantaneously occupying the region da x I(r) :
(3-16)
V = Jl-
If we introduce the velocity v' of the particles in AVt
with respect to the
mass centre by the position vl = V + v'l ,
and recall t h a t the m o m e n t u m in this frame vanishes, we can write the total surface energy PsEtot
*
= Ps K ^ l = ' » " , P(e + 5 v2)J I(r)
as follows PsEtot
= lPsV2
+
PsE,
d£ i
57
CHAPTER 3. BALANCE LAWS
where P„E = pjim E, = Urn s s ;-K>
'
f-vo
p(e + ±v'2)J
(3.17)
d£.
Hr) We finally look for the surface angular momentum which is the limit of Ml =
(r + £n)x pvJ d£ = rx Pt + nx Hr)
£pvJ d£ ,
Hr)
which, owing Hi), can be written as M t = r x Pt + psd{ n x V s , where dt represents the distance between E and the centre of mass of the particles instantaneously lying in da x I(r) along the normal n to E. From this equation and (3.15), (3.16) we deduce that M = Urn M, = rx pV .
(3.18)
Now we have to evaluate the difference
6 Urn -^M \ H » M St)
(3.19)
where Ft should be identified as psl , Pl , psEt or Afj . It is quite obvious that assumption iv) implies the vanishing of the difference (3.19) these quantities except for the last one. In fact, we obtain for this case
for all
58
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
,. hm
l-*Q
6M
l —jr-iOt
-W='."xy'('te$)"
(3.20)
The limit appearing in the second member of this equation can be easily calculated when we recall that d[ dj represents the distance (along the normal to r) E) between E and the centre of mass instantaneously belonging to da x / (I(r) as
Sd, lim6-£-=Vn-cn. /-K) St n
n
(3.21)
To conclude, we can write
fe^-W = ^.-«J-»K..
(3.22)
system with an interface. interface. 3.3 Balance equations for a system We are now in the position to formulate formulate all the balance laws for a continuous (£, ,C 2 ,E) [15]. In the sequel C system with an ari interface (Cj,C c == Cj ca U u Cc 22 ,, KV CCCC isis V ffll EE . For the passage from integral laws and any material volume and a = = V the equivalent local equations we resort to (3.1), (3.4). Conservation JL(
of Mass
pdv + pdv \/
da pPssdaj= )=
cr
PsKvN-WW—Tpf-Vs-Vxldl ^^-Cn"-JV)^-^-"E]'« a
(3.23)
59
CHAPTER 3. BALANCE LAWS
where p p and ppss are the volume and surface surface mass densities respectively. T Thus hus
we locally locallv have pp + V ■ v == 00, , + PpV-u
in CC--EE
^*jf+V + sV, -(p•sV(sPs)-2Hc Vs) -nPs2Hc -lPU} = 00 ,, n Ps - [PU] = r-[nx(C-V)p,] r.[nx(C-V)p,]
Bt ilance of
dt\
pvdv
=O 01t
(3.24)
on o nEE- -Tr
on T. onT.
Momentum
+
n • £ da +
P,Vd
pfdv +
v^ ■ Tdl + pssFdcr, Fdcr,
av +
p^[(^-c„»-^)~af-*v*'E]<«
(3.25)
veloci ty, stress tensor and body force inside C --EE where w, v, t and and /f denote denote the velocity, T and and F fields on E. Hence we write respectively; V K, T F are the corresponding fields
locally locallv -^(pu) + (V fapv) ( V --v)pvv)pv-V V-•t*- -p/f» / = = 00,, in CC--EE #c „ '
(3.26)
C/ ++nn-t]-p ■ 0 -s-P F SF~-== 0, 0, on E -- T, 1',
r - [ n x ( ( C - V ) p , V + T)] = 0 ,
By noting &"> =
= pu + pi>
on T .
60
-PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
v Up* )>-=St6-irv+Pfw St 'St KHs
vs-(v,®Psv)=vvs-(psvs)+psvs-vsv
and taking into account (3.24), we can transform (3.26) as follows
pi> = V -t + pf ,
Pi
(8V \6t
+ vs ■vsv
)
■
■
+ T)] = 0, onT.
Momentum
■jjL(\prxvdv+ ( prxvdv+
\ppssrrxV x V da da)J + er
y
rx N t da + dv
rxV[(vNNp.rxV[(v Ps
c ) pss(V n P (Vn-n-c n)
c
rx pf dv + V
+
(3.27)
-V)U + n ■t} + PsF, on E- -r,
=vs •T+[p(v-
T-[nx((C-V)PtV
Balance of Angular
inC-E
d
nxV da nxV da
T
rxn-T
dl +
rx p F da
cnn-N) n - N) ,--^-V-u^dl. ^ - Vs • „ £ ] dl.
(3.28)
da
The presence of the third integral on the left hand side is due to the considerations of previous section. Locally (3.28) implies that
j-t{r x pv) + (V • »)(r x pv) - V • (r x t) - r x pf = 0 ,
£ ( r x p,K) + V s • (V. ® r x psV) - 2Hcn(r x psV)
(3.29)
61
CHAPTER 3. BALANCE LAWS
+ P.(Vn-- cn) n x V - Vs • (r x T) - [(r x /»«){/ + r x n - f ] - r x / ) / = 0 , r •[nx [(C--V)(rx Noting that rxn
■1 =
PSV) + rx T\] == 0 .
n r x 1 , the first of these equations, can be written as
(p + pV- v)(r x v) + prx v -- V ■ ( r x
0--rxpf
== 0 ,
consequently, owing and (3.27) e't ■ xe ing to to (3.24)! we have e'' t xt e,=; = ei (3.27)!, 0 0, ,where e(3.24), and x , we is a spatial basis and e* its reciprocal basis. It is quite obvious that this relation impl:ies the symmetry oft: t=tT,
in C - E .
(3.30)
Similarly, (3.29) 2 become: 65 nPf (( St
+ +
r + v s ■(P*V.)c nnxV V ■iPfV.)- -2Hct nPs)( ,/>„) (rxV) * ■V) ++ Pspc„nx
0 a rxp rxPs6s-?-+p -£+p -asa-a ®(rxV), ®(rxV), sVfsV a a
a -rxV T+a -Txa a- r x sV T+a"-Txa s a-
+ Psp(V --cJnxV s(Vnn--c n)nxV
-- rr x = 00 .. sF = x IpvU IpvU + n-t}-rx +n-tl-rxppsl<
On the other hand a ®(rxV), ■ K ®(rxV), P: Pt>v. y. ■(a a = -V Vxaa +
and id
V ■ K '®{a '®{aaa J =-- rpps V J= s r s s
a PsVs ■a rxV,a=
! xV)) + xV)) PsVs ■ K ® ( r x P, J) a)) +py. !
v,
a
- V x ^ + r x ^ ■° )v,a
62
-PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
^ „ » * Vs + PsVs ■ VSV = - Vnn x V + PsVs ■ VSV
so that the foregoing equation becomes
r X
{P'-W
+ P V
' » ■ V ^ - V « T - P'F
~IP{V- -V)U + n-t])+
a a - T x a a = 0,
and it follows from equation (3.27)2 that aa ■ Tx an = 0 , o n E - f .
This relation is equivalent to the conditions rpOt/3
r l
T (^Oi
rpa3
n
(3.31)
which imply that T is tangential to £ and its surface components are symmetric. Finally equation (3.29)3 can be written as
rxr [nx [(C- -V) T\h=0 rXT-lnx[(C-V)p + T}j=0 sVPsV + and it is identically satisfied in view of (3.26) 3 . Balance
of
Energy ■
dt\
p(^v2
+ e)dv
+
ps(\v2
+ E)da) =
n- (t-v-h) dv
da
63
CHAPTER 3. BALANCE LAWS
v • ( T )V- -K) da+^p + \pf-vdvv dv + \v(T-V-h dcr+ F 4-J,/. s
+
s
\psF-Vda V da
a
da
V
^ ^[(V - Nc -^ cn-n i •VN)) ~- i^ y- -^ V,- ^■ ]* s^] dl. . Ps Mfti ^V2 ++ E)
(3.32) (3.32)
a
where e and ft ft are the specific internal energy and
the heat flux flux vectors
respectively in in. C C -- EE, i? E and and hfts s have have similar similar meanings meanings when when we we refer refer to to EE ---r. T.
Wee now obtain from W from (3.32) that that 2 j^p P (( ±v e))++ pp((±v ±u 2 + + e) ej V s • vu - V V■ •{«■«(t ■ v ft) - ft)- pf-v - pf ■=v 0= , 0 ^ ^ 2 + e))
iMr
,
¥
£ -2Hc + VV*S ■(p ays V £ ('. ( 2 ^ ++ EE)) )) + • (P* s( Y + +E))» - 2HcnPs( nPs(Y\V+2 E) + E)
-WSs-{T-V-h F-V-{p ■(T V ■K)s)-Ps -v -PA"
2 V- -Ipfy {\v2 + e)U e)U + n (t n-{t-v-h)}={\ v --*)]==0
T-lnx[(C-V)p r • \n x [(C-s(±V V)2Ps{\v2
+ + E) E) ++(T.V-h (T.V-s)}}ft,)] ]== 00. .
(3.33) (3.33)
It is quite easy to verify that (3.33), can be written as follows: follows:
pe-tr(t®Vv) pe-tr(t®Vv)
(3.34)
+ V-h=0. V ft =0.
By the following simple calculations which employ (3.24) 2 we find that
)y ■v.v . '».(¥'+EE)Y
61 81 TA* ( * • "
= (lv*
+
ur ■(p. \\y>
^)(t + v •ip.v.)a
-2#c 2Hc „/>. ( ^ + +Ey0*.> + E)
A¥*
-2Hc nPs) + PsVs ■Jt+P'
E)
6E 'St
64
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
2 + Pspvsvs-v{±v* a-v(}V
++
E)yui+Psvs+.%+Ps E)=- .(lV2 E)=~(lv* + E)lpU}+PaV+s.6V Ps§§
2 + E) ■ V ^ + +P PsV 3VSs-V(±V E) ,,
a a -T-V-ajaa=-V -T-V^ -T.V!a, = -a".(T.V\aVs-(T.V)=-a = -Vs.(T.V) s-T.V-a a,
- VS-(T-V) ,
.
■
.
.
.
from which we obtain 6E
6V
VE + v sv- -v T ) . Vv ■VE P>yy* P.s6t ++p > (P,*6t + p*PvSV, />4f >■ +{p°Tt+ * •■v ^ - v *s •• T) ■
- Qv2
+ E) [pU]= p&
+ PsVs ■ VtE+paF
■V
2
a 2 E)lpUl -lp(v-V)U ~{p(v-V)U + + n-tj.V-an-q-V-a -Ta-T-V -V^-^V ta- (±V + EJlpU}.
Finally, (3.33)2 can be written as a (8E -V E)-a°.T.V VSE)-fa a T Ps{jt+V s v ss■ Ps \6t +
-lp(^(v-V)2
■V
.hs a + VVs sA
+ e-E)U + n-(t-(v-V)-h)}
=0 , on ES--TT..
(3.35) (3.35)
Now, it remains to consider (3.33)3. By subtracting the relation (3.26) 3 multiplied scalarly by C from the right we deduce that r■■lnx[(C-V)p - [ n x [ {C-V)sPs {\{V-C)2 2 (±(V-C)
+ E)) ++ T-{V-C)-\]\= 0 , on T. T.(V-C)-hs}}=0,onr.
(3.36) (3.36)
65
CHAPTER 3. BALANCE LAWS
3.4 T h e reduced dissipation inequalities. Together with the balance equations we have to consider the 2nd principle of Thermodynamics in the well-known form
M 1A
psdv +
v
psSda) > -
c
+
pss
■n h ■ nda -
av
(VN-Cnn-N
uL-N
lhs-Vdl
da dl s "EJ >
(3.37)
V
rr
where s and S denote the specific entropy in C - E and E - T respectively and 0 is the absolute temperature. This inequality is equivalent to the following set of local relations
j-t{Ps)
+ ps V-w + V - ( J ) > 0 , i n C - E ,
(3.38)
jl (PJS) + V s ■ (p,SV) - 2HpsScn + V s • ( ^ ) - [psU -1 h - n] > 0 , on £ - T , r.[nx[(C-t>sS-I/y]>0,onr.
For the sake of simplicity ( see [15] for the general case) we suppose that 8 is continuous across E; then by the same procedure used in the previous section we can put system (3.38) in the form
p8is + V • h - i& • V0 > 0 , P >eJt + p. w. ■V,5 + v
K K
' e
on C - E
V6- -lpO(s- -S)U- -h n j > 0
(3.39)
on E- ■ r ,
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
66
rT-lnx[(C-V)p -[»x[(C - *t]>0 >.S-lAj>0 tS-±h
n T .. ,, oon
By eliminating V • h between (3.34) and (3.39)j as well as V s ■ hs (3.35)
a n d (3.39) 2 ,
we obtain
the two following
between
reduced
dissipation
C- E , in C E
((3.40)
inequalities -p(i> + s9) -p(ij s6) + tr(t®Vv)-±h-V8>(} tr(t®Vv)-±h-V9>0
where %l> = e-6s,
(3.41)
is the specific free-energy in C and
-ps(V
2 2 + + [p[fyv-V) [p[fyv-V)
+ Se') +
aa.T-V,a
+ il>-V]U-n-(t-(v-V)-h)]>0 il>-9]U-n-(t-(v-V)-h)]>0 ,, on o nE E - rr , ,
(3.42) (3.42)
where y = E-9S
,
(3.43)
is the specific free-energy on E and we define
A' = Jf+V.-V,A.
(3.44)
All the considerations of the previous two sections could be generalized to the case in which the bulk phases are constituted by polar continua governed by Nonlocal Mechanics [16] or by a mixture [17]. Finally, in [18] the interface is supposed to be polar to take into account adsorption effects.
67
C H A P T E R 3. BALANCE LAWS
3.5 Constitutive equations.
We
concluded
the previous section
by remarking
that
the
balance
equations we derived were not general enough to permit the description of the wide phenomenology associated to interfacial phenomena; however, they are sufficient
to predict the most common aspects of phase changes. In this
chapter we propose a set of constitutive equations which permit, together with the balance equations and j u m p conditions, to formulate reasonable boundary value problems associated with phase transitions in materials which are described by those equations. In order to assign a continuous system with an interface (C-^CjjE), we have to give the constitutive equations describing the materials in the volume regions C,, C 2 and on the interface E. As is well known, the reduced dissipation inequality
(3.40) is regarded as a restriction for the volume
constitutive equations. In the same way, the inequality
(3.42)
supplies
restrictions for the surface constitutive equations and jumps. We shall show t h a t this last inequality leads to other constitutive equations. Let us suppose that Cj and C2 are elastic continuous systems; then their constitutive equations are simply given by 1> = iKF,0),
' = 5 =
" & •
89 '
(3.45)
68
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
h = h{F,9,V6)
,
where F is the deformation gradient with respect to a reference configuration and the heat flux vector h verifies the inequality
h-Ve<0.
(3.46)
T o derive thermodynamical restrictions for surface constitutive equations, we suppose first t h a t £ is also thermoelastic, i. e.,
(3-47)
* = *(*».> "a/J.O. T- -- T(PS-aap>6) >
s = s(ps,aa/3,e),
K- = K(Ps:aa0,O,
Vfl),
where
(3.48)
hssV6<0. ■ ve < o .
Relations (3.47) are motivated by the following considerations.
The
surface quantities at a point of E depend on the geometric state of the surface since they are functions of metric tensor aag . In other words, they depend on metric
variations of £ ,
dependence
of ? ,
T
not on isometric
, etc. on
transformations.
mass density
The
is a consequence
explicit of
the
circumstance whether E can adsorb or emit mass under the same defor-
69
CHAPTER 3. BALANCE LAWS
mation. On the contrary, if the total mass m(
psda
=
p? do-
that (see (1.17)) (3.49)
Ps4a' =--P0s^°-
This equation shows that in the absence of adsorption there is a relation between ps and aan . Going back to the general case, we first note that owing to (1.46), (1.41), (3.31) we have
a<*.T.Vta
=
T°'P(Va.p--ba0Vn)
=
T"Paa0.
(3.50)
On the other hand, by adding and subtracting [ n - t - C ] to [n-f • ( « - V)] and recalling that Cs = Vs (section 3.2, in) , we obtain
in-t.(v-V)j
= lpUi-lpj(cn-Vn)
,
(3.51)
where p = - n-t-n
.
(3.52)
Consequently, introducing the specific Gibbs potential on E, <7 = ^ + - § ,
(3.53)
(3.42) can be written as
-/»,(*'+ SO') + T«V/3 2 2
V n, ), ) > + (v-V) + g- *]U}-[p](c -V) + ff-*]t/]-[p](c n- n-■ V > o0 . + IP[\ {p[\(v-
(3.54)
70
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Moreover, we have from (3.47) 1
,T,/ 9'
_- 69 89
~dfsPs
n,
. ,, 89 89
n„, ,
d%Tp a^
,89., ,89 a, +
Wte'
.
(OZK) (3.55) (3 55)
'
-
Before inserting (3.55) into (3.54) we must evaluate p's , a'aa . T h e following calculations are carried out in view of (3.44), (3.24) 2 and (3.50), (1.50)
P» == ^^ ++ V, V, ■■ V VsPs 2H 2H [pU] s s ■■VV s s++ Psc sPs== - -pspVsV Psn cn ++[pU] al3 c )) v = sa (
(3.56) (3-56)
lpUj. ■ +W
+
W i t h regard to the primed derivative of the metric tensor a, we deduce from (3.44), (2.21) and (2.13) that
2 e 2 V2 V 'c0= CnKp) = a/3= hKfi aP = ( af) ( ap-- "»'./i) = ^af} K r-((C„ n- -VJ^ )b^ap)) '.
a
(3.57) (3.57)
If we collect the results expressed by (3.55) , (3.56), (3.57) , we can transform (3.54) as follows . (o , 89\o,
, (Tal3 , „2 09
- Ps(S + %W.
+
-PS(S
+ (T
+ W)9S
(T-
+ P;
«
a/3
«„
an - - 2 P s ^
+Psd-p-s"
+ [p[5(«-^) 2 + ? - / i s r i
89
v
),„,
-2p.Q^)'ap
+ lp[^(v-V)2 + g-p,s}Ul - {(pg » a«0 _ 2 ^ ^ _ ) 6 a / 3 + [ p ] } ( C n _ yn) > 0 f ~ {(Pi §~ ^ - IPs J^)K0 + [Pl}(c„ -Vn)>0,
(3.58) (3.58)
*. = .♦,.£.
(3.59) (3-59)
where
Ps = * + Ps§f-e,
71
CHAPTER 3. BALANCE LAWS
is the specific surface chemical potential. A standard procedure can then be applied to inequality (3.58) to derive constitutive relations
j.a/3 T a/J _
_
S=-|f,
(3.60)
2|« a0 2 a«__ + Lp.s Qda • dps c0
((3.61) 3<61)
a P* s a— a
This last formula permits us to write the residual inequality in the form
\P[\ (v-V)2
+ g-ns]U]
+ [Ta\fi-
lpj[(cn-Vn)
> 0.
(3.62)
If we introduce the notation J = pU , we can regard inequality (4.17) as a function / of J
(3.63) +
, J ~ , (c n - Vn) which
has the property of taking the minimum value when the aforesaid variables vanish. Consequently, the derivatives of / with respect to these variables evaluated at J
+
= J~ = c n - V' = 0 vanish and then we have
( ?f -f -A/ OOOo+ + == °0. . (9-
(3-64)
" Ps)o - ° '
(Taa/3 6, \0-lPj)o
^-[Pl)o = 0O..
In particular, the first two expressions imply that
Mo = o •
(3.65)
72 ■
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Hence the Gibbs potential should be continuous across E at equilibrium. As is well known this is a classical result of phase equilibrium due to Gibbs. We note also that (3.64) 3 coincides with the result we derive from the balance of momentum equation (3.26)2 at equilibrium. It is quite natural to assume that if the system is far from equilibrium the first members of (3.64) are functions of J + , J ~ , (c n - Vn) : *
/*
(1/
.(vO).
22 2 I^pP ++ ((v+-V) (v+-V) «+-F ) + + (g-» p -s)^ +===//+++( (/(//+++, ,J, J-,{c - n-VJ-,(c + ((g-» s) ) + n)), n-VJ), 22
^p-(v--V) ^p-(v--V) >
(g-» (g-»)-)- " t>" - -«v/ 3y ++ (»-/*,)" ss
,,Q/3 Q / 63 o ((T T 6(/a/3 -[ P ]) 00 aa/3 / 3-[p]) (^ Q / 3 6a/3-|Pl)o 3-IP1)O
++ =f-(J ,./= / -(./ +
(3.66) (3.66)
,J-,(c ,J-,(cnn-Vn)), ".(«„-vj).
= ,,JJ~"- ,,(c„-v„)). (Mnc n --V^ „) )) ) . =: tt(J ^( . / +, ^ ++
From this constitutive equations for the jumps we also derive lg} = h(J + ,J-,(cn-V ,J-,(c n)). n-Vn)).
(3.67) (3.67)
It is evident that (3.66) and (3.67) vanish at equilibrium so that they are at least of the first degree in their arguments.
Chapter 4
PHASE EQUILIBRIUM
4.1 Phase equilibrium boundary value problems.
W e want now to formulate the boundary value problem relative to the phase equilibrium at uniform
temperature for conservative specific body
forces, t h a t is, for f — — VZ7, in the absence of specific surface forces. Starting from the local balance laws and j u m p conditions (3.24), (3.27), (3.30), (3.31), (3.35), (3.36), (3.65) the complete set of equilibrium equations can be written as:
V-t-pVU
= 0 ,
inCjUCj-E,
l ?A T; P| — = IL [*?]> s Jl i
o n E^'-- Ti
U11
rtarW, [g] = o , T
■ [n x T ] = 0 ,
73
on T ,
(4.1)
74
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where p = -n-t-n
and tf = t■ n + pn .
The boundary conditions are represented by the prescribed external pressure pe on a part dC C 5(CX U C2) = <9C and by the contact force 7 on the line dY, n SC such that
t■n = - p N ,
T-i/ E = 7 ,
on dC
on<9Sn3C,
(4.2)
where, as usual, i/£ is the unitary vector tangent to E and orthogonal to <9E (see fig.10).
fig-io Let us verify that the surface nature of T and jump condition (4.1) 5 do not permit
E to have a geometrical discontinuity line T. In fact, with the
notations of section 1.6, condition (4.1) 5 is equivalent to T + -v£ +T~
1/2 = 0 ,
onf
(4.3)
75
C H A P T E R 4. PHASE EQUILIBRIUM
whereas the spatial character of T is expressed by the condition n ■ T = 0; if we denote by « the projection of T - i / s on the plane orthogonal t o the tangent vector r to the curve T, we have u = An ~ + fin+.
It is now a very
simple exercise t o verify t h a t the conditions ( « - n ) + =(u-n)~ the
angle
between
n~and
n"""
vanishes.
This
result
imply that
means
that a t
equilibrium £ may have a t most a discontinuity line for the T -T -v-^ ■ For this reason in the sequel we suppose that T is absent. W e shall also assume, that [< s ] = 0 in (4.2)-^ This hypothesis is appropriate to the case of fluid phases and it seems a reasonable assumption if the region Cj, for instance, is occupied by an elastic body. More generally, it could be substituted by a constitutive relation [< ] = / ( p
+
,F~ ,6) .
In the next three sections we will consider the phase equilibrium of fluids, so t h a t we adopt the following constitutive equations t = - p l where / , /
denote the unit
,
T=7 J S ,
tensors in the space and on the
(4.4) interface
respectively, and j is the surface tension. As is well known the isotropy of fluids implies t h a t 1> = rl>(p,6)
,
* = *O>„a,0j,
P = p2
^=-^;=p(v>e)
where p(v,0)
'
7 = 7(^,M)
(4.5)
(4.6)
is invertible with respect to the specific volume v. From (3.53)
and (4.6)j we obtain v = ^
= v(p,d).
(4.7)
76
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
T h u s equation (4.1)-[ becomes
dg , Vp + VU = V(g + U) = 0
Finally,
the
phase
equilibrium
system
for
fluids
under
the
action
of
conservative body forces can be written as
g(p,0) + U(x) = const. ,
(4.8)
y = const. , [p] = 0 ,
h(p .*)] ==
0
with the boundary d a t a p = pe
,
on 5 C
,
fvz
- 7 ,
on 5 E n 5C .
(4.9)
4.2 Equilibrium of fluid phases separated by a planar interface in the absence of body force.
We start with the elementary case of a fluid system ( C j , C 2 , E ) which is not subjected to forces and has a planar interface. Equations (4.8) then reduce to p(x) = p=Pe,
VxeC
fo(p,»)] = 0.
(4.10)
77
CHAPTER 4. PHASE EQUILIBRIUM
According to (4,10) 2 the equilibrium value p of the pressure is uniquely determined by the condition that the Gibbs potential is continuous across the interface for any value of temperature which permits the coexistence of two phases.
T h e variance of this system is one since there is only one free
variable ( for instance the temperature). W e can say t h a t the knowledge of the constitutive equation
g(p,0)
permits, at least in principle, to derive the function p = p (9) which supplies the equilibrium pressure corresponding to any temperature when the interface is a
plane. From
(4.10)„ 5 it is also possible to derive the
differential equation whose solution is just the function p(9).
Clapeyron
In fact, if we
put
f(pW,e)
= lg(p(9),6)]
= 0,
a n d suppose t h a t equation (4.11) admits at least a solution pQ,90
(4.11)
, we have
( see (4.7)):
%
(Po-*o) = [ff](Po.*o) = M ( P o A ) # 0 ,
since in a change of state we have a j u m p in the specific volume across E. Differentiating (4.11), we obtain dp _ d0~ On the other hand we have
ldg/89] [dg/dp]-
g(p,9) = ip(p,9) + pv = ip(v(p,9),9)
l4 1/j
'
+ pv and
78
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
therefore
8q d i > . + 86 ~--89 86+lJ
p
dv drpdv. dib 89 89'~" 8v dv 89 89++ 89 89
PP
dv
89~ 89 ~
~s ' '
provided t h a t we recall (4.6) 3 and (3.45) 3 . From (4.12) and (4.7) we obtain the Clapeyron
equation
dp \{6) dp___W)_ d6-$\v{p,6)\> d6-0[v(p,9)]'
(4U) (4.13) [q l,i>
-
where \(9) = 9ls} \{9) 9{s\ is the latent
(4.14)
heat.
Before advancing further, we need some general properties of the function g(p,9)
which will be suggested by the physical considerations in the next
section.
4.3
A b r i e f r e s u m e of p h e n o m e n o l o g y of s t a t e c h a n g e s .
It is well known that in the elementary Thermodynamics [19,20,21] the set of homogeneous equilibrium states of a pure substance ^B is described by a state equation having the form
pp = p{v,9) p(v,9). .
(4.15)
79
C H A P T E R 4. PHASE EQUILIBRIUM
It can be experimentally shown that (4.15) is not defined for all positive values of v and 9. More precisely, the qualitative behaviour of surface S given by equation (4.15) is represented in fig. 11 where the forbidden states are indicated by shading. This figure shows that the body can be solid, liquid or vapour. Moreover, it can pass from one state to another by (very slow) transformations which are geometrically represented by curves on surface S. Of
particular
transformations. planes p = const.
interest
are
the
isobaric,
isovolumic
and
isothermal
They are obtained intersecting S respectively with the , v = const, and 6 = const.
Further terminology connected
with the phase changes can be deduced from fig. 11. T h e most interesting property of (4.15) is represented by the possibility to put it into the form v = v(p,6)
only in the regions of S which correspond
to equilibrium states of ^ which are in the interior of Ss
, SL
, Sy.
Along
the curves l ' - l " , 2' - 2 " , 3 ' - 3 " the function v = v(p, 8) assumes two values representing
the specific
volumes of two phases coexisting at
a given
t e m p e r a t u r e and pressure. The curves 1, 2 and 3, that are obtained projecting respectively the curves l ' - l " , Clapeyron's
2 ' - 2 " , 3 ' - 3 " on the plane p, 6 are called
curves corresponding to sublimation, melting and vaporisation.
In particular, at the point A where, the curves cross each other, three phases coexist and have assigned values of specific volumes. Another remarkable aspect we deduce from the qualitative behaviour of function
(4.15) is represented by the existence of three values
Pc,8c,vc
corresponding to the m a x i m u m of the curve 3' - 3". For these values, the liquid and vapour phases become indistinguishable (opalescence phenom-
80
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
enon). Moreover, when 9 > 9c , no increase of pressure permits to obtain the liquid phase. Then one says that the point of the space p, 9, v corresponds to the gas phase. The values pc,9c,vc
are called critical values and the isotherm
9 — 9 is the critical isotherm. This curve exhibits a horizontal inflexion at the point pc, 9C, vc.
Finally, we observe that corresponding three critical
values of p, 9, v do not exist for the liquid-solid transformation.
fig. 11
CHAPTER 4. PHASE EQUILIBRIUM
81
A first theoretical a t t e m p t to determine (4.15) was made by Van der Waals who derived the following equation by statistical considerations
p
=
where r = R/M
r6 v -bb v - b
a v V
2 '
(4.16)
is the ratio between the universal constant R of gases and
the molar mass M whereas a and 6 are two constants depending on the substance. If a = b = 0 , (4.16) reduces to the equation of perfect gas p-r9 f — v •
P=r£.
(4.17) (4.17)
T h e presence of 6 is due to the finite molecular dimensions and the term a/v reflects the molecular forces due to the cohesion. W h e n p and 6 have prescribed values, equation (4.16) is a third degree equation in the unknown v
pv3 -—{pb (pb ++rd)v r9)v2 ++av av- - ab ab== 00. .
(4.18)
T h e critical isotherm intersects the straight line p = pc in a triple point at which the aforesaid isotherm has a horizontal inflexion. This means t h a t equation (4.18) a d m i t s a triple root so that we can put it into the form
(v-vvc-f vcf = Pc Pc( = 0.° •
(4.19)
equation with (4.19) we If we put v = -~-vVcc in (4.18) and compare the resulting equation
82
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
find t h a t
vc = = 3366 ,
Pc Pc
a
-= ^ ao 2, 2762 ' 276
8 a
(4.20) (4-20)
0cc = 4 - % . c ~ 27 rb ' 27 rb
These relations permit us to express vc ,pc
'
, 9C in terms of a, b, c and vice
versa. It is possible to give (4.16) a form
independent
of the
particular
substance and then to compare the obtained equation with the experimental behaviour of (4.15). In fact, introducing the nondimensional quantities
p == p/ Pc, ' P P/Pc
Tr
== e/e - v/v e/e c,c, vV = v/vcc, ,
equation (4.16) becomes pP _=
8r 3L . —pW - 1 VV22 ' W -I
(4.21)
This equation has a form independent of the substance since it contains only numerical constants ( principle of corresponding states ); moreover,
the
isotherms we derive from (4.21) are given by the curves in fig. 12. These
curves
agree
well
with
the
experimental
behaviour
of
real
isotherms in liquid, vapour and gas phases with two exceptions. First, they enter into the forbidden region, which is bounded by Clapeyron's curve, and there is no theoretical criterion to define this region. In the following section it will be proved t h a t this problem can be solved provided that the Gibbs
83
CHAPTER 4. PHASE EQUILIBRIUM
potential of material is known. Secondly, the states defined by the triplets p,9,
v
can go into the forbidden region when any of the phases consists of
one or more separate regions ( for instance drops of liquid or bubbles of
fig. 12 vapour) having very small diameters (10
-=-1 mm).
This circumstance will be explained in the next section by supposing the interface to be able to exert a surface (isotropic) tension. In this situation, the liquid drops can be at equilibrium with their vapour which higher t h a n the pressure p
is at a pressure
corresponding to the equilibrium with plane
interface (superheated liquid). Similarly, the vapour bubbles can be a t a pressure higher than p without becoming water (see fig. 13).
84
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig. 13
4.4 Equilibrium of fluid phases separated by nonplanar interfaces.
In this section we confine ourselves to a local analysis of the boundary value problem (4.8)-(4.9). To begin the analysis, appropriate hypotheses, which are suggested by remarks of the previous section, must be satisfied by the constitutive equations of the two fluids in order that a solution exists of the posited problem. More precisely, by omitting the dependence on 9 since
CHAPTER 4. PHASE EQUILIBRIUM
85
we are interested in equilibrium configurations at uniform temperature, we shall suppose that the specific Gibbs potential g^(p) of the two phases are such that the equation
92(P)- -9\(p) == 0 , is satisfied for at least one value p of the pressure (corresponding to a plane interface). Since (see (4.7)) d ^9i2 _ _ L1 ±i on
dp 5 ^2 = 7P2^ *0 ,>
w
V Pi,P2> VPl,p2,
we can say that the equation
(4.22) (4.22)
defines implicitly the function p2 = p2(Pi)> where pl varies in a suitable interval
[a,b]3p
and p2 in another interval
[c,d] $p
. On the other
hand, by (4.8) 1 we get Fil{(x,p X,pit)
) + U(x)-c = gil(Pi)+U(x)-c l = 0. i
(4.23) 0
If we fix the value p
0
e (a, b) of the pressure p1 at x e C 1 U C2 , a
corresponding value of Cj is determined by (4.23) which now defines implicitly Fl(x°,p~)
the function pj = px(x,p~)
because d F j / d p j = 1/pj / 0 and +
= 0. Moreover, a value p e(c,d)
can be associated
with
86
p
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
e (a, 6) by equation (4.22). If we assume that
Urn p2(x,c2)
= p+ ,
xgC2 we obtain a value of the constant c 2 and then a solution p 2 (x, p + ) of the equation (4.23) for i = 2. Let us now suppose that the equation
H
= 2 [P2( a : 'y' 2 r ( ; E 'y)'P
+
)-Pl(a;'2/'2:(:r'y)'P_)] '
with the boundary condition (4.9) 2 admits only a single solution containing
z(x,y)
the point x . It remains to prove that the condition (4.8) 4 is
satisfied at every point of z(x, y). This is obvious because, owing to (4.23) the difference [jr] is constant on E ; but it vanishes at x p
_
and p
+
because of the choice of
.
To conclude, we can state that the external pressure being given, the pressure p
+
is determined at a point x e C2 ; if p ~ is the corresponding
solution of (4.22), p 2 is assigned and hence
4.5
z(x,y).
E q u i l i b r i u m of fluid p h a s e s s e p a r a t e d b y s p h e r i c a l i n t e r f a c e s .
For the sake of simplicity, the body forces will be supposed to be absent so that the equilibrium system (4.8), in the presence of spherical interfaces,
87
CHAPTER 4. PHASE EQUILIBRIUM
can be written as p- = c, (const.)
,
in Cj U C2
(4.24)
j = const. , 2T
-=- = c (const.) , [p] = c > 0 ,
fo(p)] = 0, where R is the radius of the bubble or drop and 7 is a positive function of 9 (see [22]). To analyse system (4.24), we suppose the existence of a value cM such t h a t (4.24)4
5
a d m i t one and only one solution p~(c)
C
[ O I M ] - Owing to (4.24)j , p~(c)
, p
+
, p~*~(c) , Vc e
(e) coincide with the pressure fields in
Cj and C2 , respectively, i.e. Pl(vl)~
V2(v2) = P+
V~ ,
'
so t h a t it is not necessary to distinguish between either p x or p ~ and p 2 and p+. v+
When the external pressure is given, the specific volumes v ~ = v^ , = v2 can be evaluated. In order to guarantee the existence of one and
only one solution of (4.24) 4
5
in the interval [0,Cjy] , we suppose, in
agreement with the experimental results, that the function p(v,9)
is defined
in D — (6,00) x (6^,00) b > 0 , # + > 0 and satisfies the following conditions: i) a critical value 9 of 9 exists such that, for every 9 < 9c , the function p( ■ ,9) e C (6,00) ; moreover, dp/6v
> 0 in (6,00) and
lim p(v,0) = 00 , v-*b
Inn p(v,9) v—>oo
,
(4.25)
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
ii) for every 6> G (*, 0C) , the function p(.,9)e
C1[(6,i/1((9))U
(v2(9),oo)]
Vj(#) < v2(9), and the following relation holds
Umv 1 (^) = lirn , # ) ; c
c
Moreover, dp/dv < 0 in (6, Uj(0)) U (v2(9),oo), and $p lim -5^ = 1 ^
Pi = p(v-t,6) < p2 = p(v2,9) ,
dp -=- = 0 .
From now on, the temperature 0 will be discarded in all the formulas to simplify the notations. The hypotheses i), ii) imply that the function p(v) is invertible on (6,oo) for 9 > 9C and the inverse function u(p):(0,oo)-»(&, oo) is decreasing ; similarly, when 9 < 9C , by inverting the decreasing function p(v) on (bjV-^) and (t>2,oo) respectively, we obtain two functions
v1(p):(pvoo)
-* ( 6 , ^ ) ;
v2(p):(Q,p2) -» (v 2 ,oo) .
By recalling the thermodynamical relation (4.7), two functions g^(p) and <72(p) are obtained; these are defined to within linear functions 4>^{9) of 9:
9i(p) =
«,(?) dp + 4>i .
When we take into account the relations
g=*».
d29 _0v dp2 ~dp~
_(dp\-in \dv)
< U
'
(P^PI,P2) >
89
CHAPTER 4. PHASE EQUILIBRIUM
we conclude t h a t
the functions ffj(p), p e ( p j , oo) and v2,
pe(0,p2)
are
always increasing and present an upward convexity. Moreover, since
"i(p') < « 2 (P"),
Vp'e(pi,oo),
V P "e(0,p 2 ),
at every point, the function g-^(p) has slopes less steep than those of the curve 92(P) ■ It is now possible to determine the locations of these curves on the plane (p, g) because the functions fll1(p) and g2(p)
are not completely assigned
owing to the arbitrariness of the functions <^>j and 2 of the temperature. In order to reduce this indetermination
of g-^ and g2 , together with
the
properties of <7j and g2 above derived from i) and ii) , we assume that in) for every 9 < 9C , a value p 0 e (pvp2)
exists such that
ffi(Po) = 92(P0) -
(4-26)
and furthermore lim p - * oo
g^p)
= oo ,
lim g2(p) = - o o . p —► 0
It is obvious t h a t condition (4.26) is equivalent to the existence of a solution of (4.24) with a planar interface. It is also convenient to observe that (4.26) determines the difference <j>-y-(f>2; therefore, for 9 e (#*,# c ) , <7j(p) and
g2(p)
are determined up to an arbitrary parallel translation to the <j-axis. As it will be proved in the sequel, the remaining arbitrariness of g^ and g2 does not affect the physical results. Finally, we observe that (4.26) and the previous
90
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
considerations imply that (4.27) (4.27)
9922(P)<9I(P), (P) < 9i(p) ,
Pe{pvp0), pe(PvPo)
92(P) >9\{P) (p) > SI(P) .<
P£(P p e ( P0,P o 2' )P 2 )■-
>
In order to prove an existence and uniqueness theorem for the system (4.24), we start with the relations gi =
(4.28) (4-28)
g 2 = ^922(P ( ^2)2 ) '.
go = ffi(Po) ffl(Po) = 92(Po) (P0) >.
„ - „ „ f of „ f fthe i , „ following f ^ n „ „iTiriff rtheorem l-i A r t n # \ r v i and the proof
THEOREM
1. — The functions
are
g-, and g2
both invertihle
on the
interval [gi»gg]; moreover,<> on g 0 ] 92~[ (1g{g)-9r ) - ~9i (s) H
1
is positive,
decreasing
P R O O F ..— --The The
and vanishes at gQ. function
^<7j(p) ( p ) increases
in
[pj,oo) [p^oo)
and
{p) p j , o o ) . Similarly, the function function g22(p) invertible on [p 1 , p 0 ]j C [ [pj,oo). [Pi-Po'
then
it
is
increases on
i l is also invertible on [[p-pPg] (0,p 22 ]- Since ;p°i1 < p0 < P ^ P Q ] .■ By (4.27), (4.27), (4.28), < pP2,it < Po 2.
we have have g2{[p9v2p([P - [g2(P2)>So] 0)] we VPQ)]~= L 9 2 ( P 2 ) .Sol both
g^~ J (g) (g) functions g-^~
{d/dg)g-\g) (d/dg)g,r\g)
,,
cc
[gi-Sol- Therefore, Therefore, on interval [[g [gi-golon the interval g ia ,,g g 00 ]] g2~ (g) exist and are increasing because
= = l/vi 1/*,i > 0. Finally, from v
d " I " - 2
ff2_1(gi)
= P*
an<
^
recall that
ffi-1(gi)
91
CHAPTER 4. PHASE EQUILIBRIUM
= p 1 , t h e previous result allows us to say that the difference g^ (g) S' 1 _1 (g) a t t a i n s its m i n i m u m p* - p1 a t gv decreases in the interval [g 1 ,g 0 ] and vanishes at g 0 . By setting
*™ = ^ r '
(4 29)
"
Pv'Pl we are now in the position to prove the following theorem: T H E O R E M 2. —If the temperature
9 is fixed in the interval (O^.,0C) for
every value of c e (0, pv - Pj) the system (4.24) has one and only one R
(Pi> PV' )<
where
e
e
Pi [Pi ' Po\ • Pv [P* > Po)
and R
solution
e (Rm, oo) •
P R O O F . — Owing to the previous theorem, the equation
g2-1(g)-gr1(l)
(4.30)
= c,
has one a n d only one solution g if and only if c e ( O J P ^ - P J ) . On the other hand, 2 (Pl,pv)
and g^
being invertible on [gj,g 0 ] ,
one and only one pair
corresponds to every solution g of (4.30) such that
9r\g)
= Pl,
F u r t h e r m o r e (4.30), (4.31) imply pv-pt
9i\l)
= Pv
(4-31)
= c > 0. Finally t o every e g
(0,
p * - p,) it can be associated a radius R for the spherical interface given by R = 2y/c which belongs to t h e interval [Rm, oo) since c = p * - p j is t h e greatest value of c (see fig. 14).
92
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig. 14
It is clear t h a t in a specific problem we have to take boundary conditions into account. If we suppose that the vapour is in the interior of spheres which are contained in the liquid ( bubbles), it follows from theorem 2 that T H E O R E M 3 . — If the temperature liquid
at uniform
pressure
Pi&[Pi,PQ)
6 is fixed in the interval is at equilibrium
bubbles if and only if the vapour is at a suitable pressure is contained
in bubbles of a fixed radius R e [i? m ,oo) .
(0^,6
), a
with its
vapour
p e [P*,PQ)
and it
93
CHAPTER 4. PHASE EQUILIBRIUM
It is worthwhile to observe that, when the functions p1(i<), P2iv)i 9\{p)i g2(p) are assigned (these last two being defined to within a constant), it is possible to evaluate the values of pv,R,vl,vv
for every pt e [PijPn]'
THEOREM A . — If the temperature is given in the interval (9^,9 ) , a vapour at a uniform pressure pv e [p 0 , p 2 ] is at equilibrium with drops of radius R of its liquid if and only if liquid has a suitable pressure p ( e
[PQ:P*]
and R e [2y/(p* - p 2 ), oo]. Theorems 3 and 4 lead to the same results that Serrin proved in [23] starting from the Korteweg theory of capillarity [24].
4.6 Equilibrium between isotropic solid and fluid phases.
In this section we consider an isotropic elastic solid which occupies the region C 5 and is at equilibrium with its melt or vapour in the region C^ [25] . We suppose that the body force is negligible, the condition [<SJ = 0 is satisfied (see section 4.1) and the external pressure p e is uniform.
From
(4.1), (4.2) we have V * = 0,
inC5-E,
Pp = const.= p e ,
in Cp - S ,
T?g = 0 ,
onS,
Ta\0
Pe-Ps'
=
(4.32)
94
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
9F where ps = -n-t-n
~ 9s = ° '
i and t-n
= -psn,
(4.33)
owing to the hypothesis [< s ] = 0 and tp — — peI. In the traction boundary value problem for an isotropic elastic solid Cs we specify a natural reference configuration C 0 , in which the mass density p0 of the solid is uniform,
and write (4.32)j and (4.33) in terms of the
Lagrangian coordinates. A displacement u(X),Xe£0,
which is a solution of
this problem, supplies the equilibrium configuration C s . In the case we are considering, the reference configuration CQ cannot be assigned completely. In fact, we can assume that C 0 is a homogeneous natural state at a uniform pressure p 0 , where p0 is of the order of the equilibrium pressure of the solid in the presence of its vapour or melt at a given temperature. However, we can not specify <9C0 because neither the total mass of the solid nor the region it occupies are known. On the contrary, as we shall prove, $C 0 is determined by (4.32) 3 4 . If we suppose that
C0 -* Cs is an infinitesimal deformation, the
stress tensor can be written ( see, [26], p. 249) as
t = t0 + Ht0 + t0HT - trHt0
+ C-E .
where H = Vw is the displacement gradient, E is the infinitesimal deforma tion tensor and C is the
infinitesimal elasticity tensor. By recalling that
10 = - PQI and the solid is homogeneous and isotropic in C 0 , we have C • E = XtrE + 2pE and the previous relation becomes
95
CHAPTER 4. PHASE EQUILIBRIUM
tt=-p = 0-I pQI + trE( trE(p + X)I A)/ + + 2(AI 2(/i -- P Po0 + P00)E )# •.
(4.34)
ion we deduce that From this relation trt + 3p0 itrEr^ = 5-r—S- , + 3A ++ 2/z ~ pp 00 + 2p ' 3A + 2p
1
0
so that we have
we have Zr*+ +Zp3p 0 l1 trt P \ti I/ 7 - ((v i| AA\) 0 iJ [ +Pv (Ppo + )) [* P o + E [t+PoI {Po X) 3+ 2/z p + 33A ^ -=2 2i^pV) ( / , - p 0 ) " -" u 2 ^] - ' PPo o0 + A+ +% -
(4.35)
(4 35)
'
On the other hand, the specific free energy tf)g(H) up to terms of the second order is represented by the following expression:
i>s(H) = ^s(I)-^trE,
(4.36)
where V's(-0 is a constant owing to the homogeneity of CQ. Therefore, (4.32) 5 can be written as follows
iPs(I)-P<±trE
+^ ( l
+
trE) = gF(pe),
on E .
But we have also Pg = - n -1 -n = p 0 + terms of the I s order. Hence
MI)+jPj
= 9F(pe).
We recall t h a t rpg(I) is undetermined and gp{pe)
(4.37)
is known up to an additive
96
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
constant
(when the temperature is given). Moreover the quantity ps
expressed by the difference
P(,[gp(pe) - ips(l)]
is
which contains only one
additive arbitrary constant. This means that (4.37) permits to express ps as a function of p e when we experimentally know the value of ps corresponding to a particular value of p e . If we denote by p a plane interface,
the equilibrium pressure with
we deduce p s = p~e from (4.32) 4 . In conclusion (4.37) can
be p u t into the form PS~Pr
= Po[9F(Pe)-9F(Pe)]
>
on S
(4-38)
>
which does not contain arbitrary constants and shows t h a t ps is uniform on the whole boundary <9C5. -Psn
According to (4.33), t-n=
on E is a uniform
pressure and
therefore a solution of (4.32)^ is represented by the uniform stress tensor t = - p$I. This solution is acceptable because (4.35) shows that E is uniform and consequently satisfies the compatibility conditions V x V x E = 0. By taking all these results into account , (4.35) yields the form
£ _
Ps-Po p 0 + 3A + 2/z
f
Pol9F(Pe)-9F(Pe)] p0 + 3A +
+ Pe^
2/J
I = f(Pe)
I-
(4.39)
It is easy to verify that, up to a rigid displacement, (4.39) implies that
u = f(Pe)X,
XeC0.
(4.40)
97
CHAPTER 4. PHASE EQUILIBRIUM
On the other hand, by (3.64)1, gs = <7^(pe) = / i s ( a c*3' Ps)
anc
^
we can
'nen
express ps as a function of aQo so that the equation (4.32)4 becomes r""(a^)6Q/J=pe-Ps, where ps in turn
(4.41)
is a function of p e owing to (4.38). Consequently, the
pressure pe being given, the relations (4.32) 3 and (4.41) represent a system of three equations in the unknowns aao, bag. When we add the Gauss and Codazzi-Mainardi equations (1.28), (1.29), we obtain a system which, in principle, is able to determine the interface r = r(ua)
up to a rigid
displacement. In conclusion, recalling that the actual position x — X + u, we deduce the boundary r0(ua) of C0 corresponding to a given pressure as
r (
» "° ) = TT7TO'
4.7
<4 42
' >
Variational formulation of phase equilibrium.
In this section the equilibrium of a solid phase in the presence of its melt or vapour is discussed from a variational view point. In other words, we do not use the equilibrium system (4.1) but resort to the Gibbs principle according to which the equilibrium configurations are extrema of the total free energy with respect to variations at constant volume or of the total
98
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Gibbs potential with respect to variations at constant pressure [25, 27]. W i t h the notations of previous section, let
\p =
pFipF
pgi's dv +
c5
dv +
p^s
da ,
(4.43)
cF
be the total free energy of the system where ipF = ipp(pp)
, ipp — ipp(F)
,
$ = ^ ' ( p ( T ! i a g ) and E is a regular surface. This last hypothesis excludes that the following considerations can be applied to crystal planes. Moreover, we suppose that the solid and the interface are isotropic so that xps depends on the principal invariants of B = FF
and $ on the principal invariants of
aa0- The supposed isotropy of the solid phase is equivalent to require that the crystal structure does not manifest itself at a macroscopic level. This is possible if the dimension of Cs are large compared with those of the crystals which constitute the solid and if the crystals are oriented randomly. Let us consider the set of functions K = {Jfc:E —► k(E), regular , Jb(<9£) = <9E and conserving the surface mass}; R = {ps:i:—>ft
+
, regular}.
Moreover, we will denote by h :C —► C , C = Cs U C^ , a mapping such t h a t i) h is regular in C - E and exhibits finite discontinuities together with its derivatives across E; it) h.fo
: C s —► M ^ s ) ' ^ | C
' —* M ^ F ) '
are
diffeomorphisms and
conserve locally the mass. If we denote by H the set of these mappings, the quantity defined by (4.43) will be regarded as a functional over the set H x K x R which can be normed in an obvious way.
CHAPTER 4. PHASE EQUILIBRIUM
99
The Gibbs principle states that (he, ke, pes) is an equilibrium configuration if ^(h,k,ps)
has a conditional extremum with respect to all the arguments
h, k, p s satisfying the global constraint tp = tp =
pps sdv+ dv + r>5 C
Pp pFdv+ dv + pps sda-M — 0Q, da - M — , y rZ >F £
(4.44) (4.44)
M being the total mass of the system. As is well known, the extrema of (4.43) under the condition (4.44) coincide with the extrema of the functional *+ $ =y + Av?, \tp ,
(4.45)
where A is a constant. In order to evaluate the Frechet differential d<& of <&, we start by observing that the variations of i>s(F)dv' dv's^ss(F)
dtf 55 = rf* = C
5
F pPs^s( srps(F)
dv
^S
pss[iP [ii,ss(F') (F')-4> - ^sS(F)]dv(F)} dv - | ps^4> -n)Skn) da da s(6h s(Sh Sn Sn-8k
= =
E
C c?5
/ 6h dK, dv 6h Sn-6kn)6k PSips{6h da• . Ps PS -Qp-FdF jL./jL i, j3 dv ~ Ps^s( Sn - n) dcr
= = i>
e5
%±j
"
Ey.
By applying the Gauss theorem to the first integral and recalling that 6h = 0 on 9C , we obtain
100
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig-15
d ^ dF 9F,LL
di>s - L *W>« dada■ 9FiLiL **) 9F **) d p ' *S= iv.IPS-IP-FjLN^hs.da\(ps®pLFjL) d * ss
6h 6h
Si Si
dv dv
1
6hSidv
6kn) da da- . ~- J PS^s( Ps^s(6h6hSn s,, ---**„) E
Now we introduce the quantities t1 = Ps Ps^¥FT'
~
dF * ps — -n-t-n
' ,
t3 = t-n+ psn , t-n+psn<
iin n CC
s5 on E
(4.46) (4-46)
(ts-n === 0) 0) (*.••
ffs = ^+jj§» which respectively define the stress tensor t in C 5 , the pressure ps on E, the tangential stress and the specific Gibbs potential on E (see (3.53)). We remark that gs is meaningless in the volume. It is quite easy to verify that dWg can be written as d*s s == dtf
V - i - 5V't>6hdv-\p ft^\ps9 - <5fcn) da S(^Sn sgs{6h Sn--6kJd
E E
CHAPTER 4. PHASE EQUILIBRIUM
101
■ (Sh Sks) dcr - ps ps8k8k + ts-(6h + + ts t■ 6k6k s-6k s s-)das ■ n ndada s sdada. . vE
V. E
(4.47) (4.47)
m E
Similarly it can be proved that
d^F= =
» Vp ■ F8h Sk n) Vp-8h dv+ F dv + p F g F (6h Fn-6k PF 9F (6hFnn) E CF
p 8k 6knn da,
d<7 da+
+
(4.48)
E
where we have defined di>F P=J^, P = dPF>
(4.49)
iinn CCF
F + jp, 9gFF = ^^F -jTF-'
on o nEE ..
In order to evaluate the variation d'$a of the third integral in (4.43), we recall (see (1.17)) that da = -\[S du du ; consequently, 8{da) o(aa) = = k■-6aoadaaa. . Moreover, owing to (2.13), (2.14), (2.15) , we have af3 8{da) = i aaf} 6k 8ka/i a0 da ,
(4.50)
8a 8kapaj3 ., af) =
4 51 ((4.51) - )
where Sk
6k afi ap
= 6k +6kSk .n = 6k c;a.pd + 0;p.aa-2b ~ 2ba^k af}nSk
(4.52) (4.52)
•
According to hypothesis i), the surface variation 8k conserves locally the mass and then from psda = const, we obtain af, 6p.= -%P,a 6kalifi.6ka0. 6p,= -%PSf .
,
,i
,i
.
.,
(4.53)
When we bear in mind that there is another mass variation Sps due to the J,.
102
-PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
adsorption on the surface, (4.50)-(4.52) leads to the result
d*„ =
P's *(P>*fi)
d
d
°'- \ps nps^afi)
°
(Ps ~ \ Ps^K? + « * + §f- ( 4 ^"^a/3 + *P.)
+ a^-^](i+^M dff -
Ps*S
d(T
+6k i-htiiraaP+w-H'k°ie *«a/3
26
«*"»>d
+ ( +
I * "'^"-* r
By applying the Gauss theorem and introducing the notations 2 d^naP^0n
rpafi _
3*
(4.54)
^=4(^}' which define the surface stress tensor and the surface chemical potential, we have
d*s = - J T $ 6ka/3 aV- J V ^ 6knda- Pjp. da . E
(4.55)
E
It is now easy to prove that
Ps(6hSn-6kn)d
dtp = -
i:
PF (6hFn-6kn)
+ i:
d(T
5ps da .
+ i
(4.56)
103
C H A P T E R 4. PHASE EQUILIBRIUM
By collecting all these results, the condition d<3> = 0 can be written as
V • t ■ 6hs dv + ^S
Psi9s + x)(^Sn
Wp-8hFdv-
~ ^n)
da
^F
PF(9F
+ X)(6hFn
~ 6kn)
d u
(fif + \)6ps
+
da +
ts • (5hs - 6ks) da E
{T?l-
The
arbitrariness
£ ) da-
+ T"&bap)
(ps-p
6knda = 0
of the variations 6h, 6k leads us again
to the
phase
equilibrium system (4.32), (4.33). We conclude this section with the following remarks. i) When "t depends on a = det(aan)
(fluid interface) and the mass is
locally conserved in every deformation of E ( that is E does not adsorb), then the relation psda = const., conclusion that a = c/p2s
which implies ps~{a = const. = c, leads to the . Consequently, by putting ^!{ps, a) = $ ( p s ) and
recalling (4.54)j, we have
j.a/3 _
But
~P2s
<9# aa(3 + 2p 5 * 0 * aa(3 + 2p d± da _ s s da daa0~ - P I dPs da dp.
aa"*3
-o2(d±. 2c -da-)" a/3 ~P* d* = d¥ dl! ®Ps ®Ps da
da Ps
d
=
d* dPs
2c d
so t h a t na/3
Tap = 7 ( P >,«/3 -
2 . 3 * . „a/3
-P
(4.57)
104
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
This relation shows that the surface stress tensor is isotropic and the surface tension j depends only on p3 . Particularly, if \P = a ^ a , that is if E is a soap-like membrane, we have j = pg^ = const. const.
and the surface free energy is proportional to the area of E. it) Let us suppose that both the solid and liquid phases are at uniform pressures pg and p respectively. Then it is easy to verify that the equilibrium configurations of the surface E correspond to those closed surfaces which bound a volume having a fixed measure V and are extrema of the total surface energy provided that the surface mass is conserved. In fact, let us look for the extremal surfaces of the functional
$ =
Pa^iPs^afi) da < X 1 J
verifying the the constraint tp = \dv - V = 0 . tp = \dv - V = 0 . Owing to (4.55), (4.56), these surfaces satisfy the condition Owing to (4.55), (4.56), these surfaces satisfy the condition -
T?jJ6kadv-
(T°%,-A)M n <7 = 0,
- 33T?j*6kade-^T°>\fi3-X)6knda
= 0,
£ E which leads to (4.32) 4 whenever the Lagrange multiplier A is identified with which leads to (4.32) 4 whenever the Lagrange multiplier A is identified with the difference p - Pg ■ the difference p- ps •
Chapter 5
STATIONARY AND TIME DEPENDENT PROBLEMS
5.1 A stationary problem and its nondimensional analysis.
In the previous chapters we have proposed a rather general model of a continuous system with an interface which can be applied to study the phase equilibrium. T h e investigation of stationary and is much more difficult.
time dependent phenomena
Therefore, we suppose in this chapter that
the
interface reduces to a pure moving surface of discontinuity for the volume fields. W e start with a model which is appropriate to describe the stationary continuous casting of an isotropic and incompressible solid. To be more precise, let us assume t h a t (C s , Cj, E) is a system with an interface which is formed by solid and liquid phases occupying respectively the regions C s and C,. Both materials are incompressible and inviscid and subjected to the
force
of gravity g. Moreover, we suppose that all fields are stationary, the interface
105
106
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
E is a surface of discontinuity at rest (c = 0) and finally the temperature 6 is continuous across E. Consequently, we have the following constitutive equations for the stress tensors t and specific internal energies e
ta=-PaI + t (1,11,0), t^-pfl, e s = e s (7,//,(?),
(5.1)
e, = e,(0),
where I, II are the first two principal strain invariants of the solid which deforms from a reference configuration C +s to Cs. Inserting these assumptions into (3.24), (3.27), (3.30), (3.31), (3.34), (3.35), (3.67) we obtain the following equations for the system we propose to discuss V-w = 0,
inC-E,
(5.2)
pv Vv = V • ( + pg, pv-Ve = t-Vv + IpvJ = 0,
V-(k(9)V6), on E,
(5.3)
[ p W n - t - n ] = 0, 2
[p ( i v + e) vn - v ■ t ■ n - kV6 ■ n] = 0, l±v2 + g} = 0, where k is the thermal conductivity. We have denoted fields which refer to different phases with the same symbol. This set of equations, though it is highly simplified as compared to the general system which involves a material interface, is still rather difficult to handle. Therefore, since we
107
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
intend to discuss the continuous casting, we suppose t h a t e, = e.(6),
(5.4)
and the velocity fields in both phases are uniform, that is vs = vsu,
vt = vp,
(5.5)
where u is a unit vector along the z-axis. Then the system (5.2), (5.3) becomes V t = pg,
in C,
p v c(0) 9,z =
PsVsn = PlVln J Plv lv} V lifJllnuu-[n-t} lnu-l"-H
[ i v22 + e\Pl Plvln ln
[l,2 +
hypotheses
considering; further
on
the
V(k(9)V0),
.
o n
E
(5-7)
>
= ~ 0, "'
- [vn • *t ]] • « -- [JfeVff] [JfeVff] ■■ nn= =0,0, V)W
j_n^n_g=0.
where c(9) = e'(9) is the specific heat. drastic
(5.6)
physical
This set is the result of the very properties
of
the
system
we
are
simplifying hypotheses would destroy its capacity to
describe the continuous casting. However not all terms appearing in (5.6), (5.7) have the same weight as could be recognized by a nondimensional analysis. In order to express (5.6), (5.7) in terms of nondimensional quantities, let us introduce certain scaling quantities for velocity, pressure and temperature (V, P, 0 ) a n d two length scales L and /. We calculate the derivatives of the t e m p e r a t u r e field with respect to z by scaling with the first length whereas variations of 9 with respect to x and y coordinates by the second one. In such
108
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
a way, the following analysis will be applicable to the case in which the system possesses dimensions of different order along the z axis and x, y axes, but comparable thermal variations along all the axes provided that the displacements are of the order of the ratio L/l. We treat the fluid velocity as scaling velocity U, whereas P and 0 are respectively identified as the atmospheric pressure and melting temperature. Moreover, if c = C/(0), k = &/(©) denote the specific heat and the thermal conductivity of liquid at temperature 0 we assume that "I < Ti c
? 0
(5
'
-8)
i.e., the specific kinetic energy is much smaller than the specific energy associated with the pressure and the latter, in turn, is much smaller than the specific internal energy. By denoting the nondimensional quantities with the same symbols of dimensional quantities, it is easy to verify that
system
(5.6), (5.7) can be written as in C,
^ t = p-p9,
(5.9)
1}c{e) 9>* = J^L ( f <*'••)■• + f <M'»)'y + (**»«)»*)' on E
Pvsn = vlw (vf-l)u—£j
»
[«.*] = 0, Plvl
[i« 2 + i #*i„=-- £ 5 [n-*]-i. —^TT ( 7 ^'*'' + M
+W
'»«] •» = 0.
(5-10)
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
109
[ l , M f *]_-£, («^3 + p l)as o, where /z = pjp\
and »', j are the unit vectors along the axes x, y. When we
take (5.8) into account, the previous system reduces to the following
V t = 0,
5*} c{9) e,z = a(^
in C,
(5.11)
{k9,J,x + £ (kd,y),y + (k9J,z),
l">»n = vlm
on S
'
(5-12)
[n.*J = 0, X(9)vln = -0[^k(0,xi
+ O,yj) + ke,xu]-n,
M = o, where we have introduced the notations: _
A *; PlVjcL
8a _- -a
Here A(0) is the latent heat. Equation
(5.12)4 supplies the melting
temperature on the interface which is independent of the velocity and pressure. This, under the hypotheses (5.8), is in full agreement with the experimental results. If z = ip(x,y) is the equation of the interface, the vector (
+ ^)1/2
110
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
and (5.12) 3 leads to the equation
\(6)vtu
■n = -
2
[ k k(9,J + 6,yj) + ke,2uj-(V
(5.13)
Pi T h e geometry of the problem we are considering is represented in fig. 14. On the section S (z = 0) of the cylinder Sx[0,L],
the pressure pe and
the
uniform entrance velocity VfU of the melt solid are given. Then equations (5.11)]^ and (5.12) 1
2
imply that t=
velocity of solid is v, = vj^i.
- peI
everywhere and the extraction
It remains to determine the temperature field in
S x [0, L] as well as the equation z(x, y) of the interface E. We are led to the following boundary value problem in the unknowns 9(x,y,z)
fig.16
, (p(x,y) :
111
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
V(x,j/,z)eSx[0,L]
-E,
^ } c{9) 9,z = a(^-
(k6,x),x
+£
(k0,y),y
+ (**„)„) ,
(5-14)
V(x,y,z)eS,
X(9)Vlu ■ n =
0(^,0) = ^,
L
[ i *(0 >jr i + 0 j ) + *0, z u] • (V
9(x,y,L) = 92,
kj^=f(x,y,z),
on
fi,
(5.15)
(5.16)
where # j > 0 , # 2 < ®i © ' s the melting temperature and $7 is the lateral boundary of S x [0, L], In [28] J. Rodrigues supplied a variational formulation as well as a weak existence theorem for the boundary value problem (5.14)
(5.16) for the case
L = I. In [29], L. Faria fe J. Rodrigues analyzed the aforesaid problem from the numerical point of view by assuming a rotational symmetry around the z axis. By supposing that the heat extraction from the lateral boundary Q. is represented by the formula / = a[9 - h(z)], they obtain the profiles in fig. 17. It appears that the m a x i m u m depth of the surface with respect to the plane x, y is affected by both the extraction velocity vs and the lateral cooling.
112
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
F,=0.5
V. = 0.7
fig. 17
5.2 On the approximate evolution of solid-liquid state changes.
The previous section was, in a way, dictated by the analysis of a stationary problem which describes satisfactorily the important process of continuous casting of metals. Now we want to analyze dynamical evolutions of state changes under the following hypotheses: i) the interface is a moving discontinuity surface for the bulk fields;
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
113
ii) the problems are one-dimensional. In this section we start with the solid-liquid
state changes. Let C be a
system formed by two phase C s and Cj of the same substance in the solid and liquid states, respectively. These phases fill the layers 0 < x < s(t) s(t) < x < L(t),
and
where s(t) denotes the location of the plane interface £ which
satisfies the condition i); moreover all the bulk fields are functions of the spatial variable x and of the temporal variable t. The solid in the region C s can
always be supposed
to be at
rest. Taking into account all
these
conditions the balance equations and j u m p conditions (3.24), (3.27), (3.30), (3.31), (3.35), (3.36), (3.65) can be written as
Pscse,t
ie[0,s(*)), *e[0,oo),
= kse,xx,
pH + (pv),x
= 0,
te[0,oo),
xe(s(t),L(t)],
(5.17)
(5.18)
p{vn + vv>x) = -P>x ' p(e,t + v e , J = -pv,x
,
p(v-s)=-pss
+ k6,xx
x = s(t) ,
,
ie[0,oo)
(5.19)
[v]J + [p] = 0, 2
[%{v-s)
+ e\J + [p(v-8)]
[i(t;-s)
where p3,cs,
2
=
[keim\.
+ ^ + | ] = 0,
k^ are three constants which represent respectively density,
specific heat and heat conductivity of C s and 0(x, t) is the temperature field; moreover,
to simplify the notations we have not affixed the subscript / to all
114
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
the quantities which refer to the liquid; finally, cn = s and (5.19) 4 is assumed to be valid also in dynamical conditions ( see (3.67) ). Equations (5.17)-(5.19) are valid also if the layer s(t) < x < L(t) is filled with vapour. We postpone the analysis of this case to the next section; at present we can further simplify
(5.17)-(5.19) by recalling the incompressi-
bility of liquid phase. More precisely, we assume that
p = const. ,
e = e{9) = c9+ eQ .
(5.20)
From (5.18) 1 and (5.20)j we at once deduce that v depends only on t; this conclusion, by taking into account (5.19)}, leads to
v = as ,
It is
where a = —p—- .
(5.21)
easy to verify that the boundary value problem (5.17)-(5.19) now
becomes 0.i = < » A « .
aps=-p,x,
xe[0,s(t)),
x e (s(t), L(t)] , Q,t + ase,x
lpj = apss2,
te[0,oo),
=
,
(5.23)
a6,xx,
x = s{t),
Pj e l* +^ a2pj3-aps
te[0,oo)
(5.22)
=
te[0,oo)
-mx].
(5.24)
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
M ♦l where a, = ks/pscs
as2- 2 -
and a = k/pc
P = "aP7 =
115
o,
are the thermal diffusibilities of solid and
fluid phases, respectively. We explicitly remark that the incompressibility of the liquid in C( imposes that the boundary L(t) moves with the velocity v of the particles of C(, i. e.,
L(t) = v(t) L(t) v{t) = as(t) as(t) ,
(5.25)
and therefore the motion of the free boundary of C; is determined by the evolution of the interface. Usually the densities of two phases have almost the same values so that a
remain
almost
the same during
the phase change.
The
boundary value problem (5.22)-(5.25), although highly simplified, remains quite difficult
to solve. However, by a nondimensional analysis we can
recognize t h a t not all the terms in equations (5.22)-(5.25) have the same weight. Let X be a length comparable with the dimension of the system C; as is well known, the quantity T = X / a conduction
phenomena.
denotes a time which is characteristic of
This means that
the interface
supposed to be comparable with the rate V = X/T.
velocity
can
be
Finally, let us prescribe
the initial and boundary d a t a for the temperature and let 0 be the difference between their m a x i m u m and minimum. By introducing the nondimensional variables T*
x _
X
X'
x
,
1
yi )' T
P
P
=
y~P'' >
* 0
V:'
* -— 6e— -' _
(5.26) (5-26)
116
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
s =
F'
e =
^e' * =7e'
where QM is the melting temperature at ordinary pressure, we obtain the following nondimensional system in which we have again employed the same symbols for nondimensional fields:
0> = 0,„ , ».xx = «,„ » p,xx= = --aAfis, aA/zs ,
t e[0,oo) , *e[0,oo),
«e[0,*(*)). «e[0,*(*)).
<e[0,oo) , *e[0,oo),
xe(s(t),L(t)], a: g (*(*), 1(01 i
ff,t + a«fi,a. = atf,a!ir
?
6,t + as6,x = a9,xx
,
[pl^a^s2, {pj-aAs2,
x = s(t), x = s(t),
*e[0,oo), <e[0,oo),
(5.27) (5.28)
(5.29) (5.29)
[e]s+Ia3s3-a|ps=(0,x)+-k(0,x)- , [e]s+Ia3s3-a|ps=(0,x)+-k(0,x)- ,
[ V - ] + i a 2 B s 2 - a | p = 0, W}+±a2Bs2-a%p
= 0,
where where
kk -=- * -
u-2**=£' »=f ,
f' k=A,
s
A-P*V2 A — ■"■
~
Pp
'i
aa -=- 2 a =
£' A,
5(5.30) ((5.30) -3°)
cs6'
and ( ) + denotes the limit of ( ) on the interface when we approach from the liquid phase. A similar meaning is attributed to ( ) _ . Tables 1 and 2
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
117
contain numerical values of the physical quantities we are considering and show that the numbers A, B and B/A are quite negligible with respect to a .
6.91
2.91
8.66
2.33
s
8.62
4.81
30.93
1
8.36
5.44
30.93
2.55
11.39
24.19
1
2.38
10.47
24.19
S
0.91
19.26
0.221
1
1.00
41.86
0.055
Aluminium
Water
k/io" erg/cms. D t
6.90
1 _
c/10* erg/8 D K
7.36
Iron
Copper
P g/ cm
0M
lat heat /10 =T8'8
1808
2.72
1356
2.13
930
3.93
273
3.35
Table 1
a
k
a
\
Iron
-0.067
0 80
0.68
2.18
2.1-10 -n
2.6-10-'6
n-io-4
Copper
-0.031
1.00
0.91
3.27
47-10" 10
8.5-10"14
1.8-10"4
Aluminum
-0.071
1.00
1.17
3.72
17-10" 10
6610" 1 4
38-10""
Water
0087
025
on
6.37
14-10" 14
3.0-10""
21-10" 3
A
B
B/A
Table 2
The numerical results contained in tables 1 and 2 permit us to neglect the terms containing the factors A,B or A/B
so that the system (4.27)-(4.29)
118
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
reduces to the following set of equations:
*.* = * . * » .
x e [ 0 , «(*)),
p = p(t) = Pe,
<e[0,oo), *e[0,oo),
xe(s(t),L(t)],
(5.31) (5.32)
d,t + as6,x = a6,xx , [p] = 0,
*€[0,oo],
x = s(t),
(5.33)
,
[e]i=(O,x)--k(0,t)+
W = o, where pe is the external pressure. We thus conclude that the pressure field is uniform and equal to pe; moreover, the condition [i/>] = 0 supplies the temperature on the interface since the specific free energy depends only on the temperature in both phases. Introducing the nondimensional latent heat A = [ e ] / c s 0 we arrive at the final formulation of the boundary value problem relative to solid-liquid state changes
»,* = « , „ ,
xe[0,a(0),
6,t + as6,x = &0,xx,
*e[0,oo), ie[0,oo),
xe(s(t),L{t)],
6=0,
x = s(t) ,
Ai = ( * , „ ) " - k ( 0 , J + , 9(x,0) = 9o{x),
xe [0,1(0)],
6(0,t) = 01(t),
«(0) = * OI
9(L(t),t) =
e2(t).
(5.34)
119
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
Carslaw and Jaeger determined a solution of the system (5.34) equipped with initial d a t a on the whole real line K, by employing the similarity method (see [30], p.209); C h a m b r e [31] found an exact solution on -ft for the state change of a viscous fluid taking into account the influence of convective phenomena. In [32] a solution of (5.34) is found in a semi-bounded domain illustrates the motion of the interface s(t) and the free boundary
which
L(t).
5.3 O n the approximate evolution of liquid-vapour state changes.
T h e liquid-vapour state change is a very complex phenomenon which is produced by the combination of conduction and convection. Inside the liquid, vapour bubbles form, they can unite with each other and give rise to larger bubbles. These reach the free surface where they release the vapour contained within. However, when the external temperature is only a few
degrees
different from t h a t of evaporation, the phenomenon takes place with simpler modalities. Let us consider, in a rigid container with a freely moving piston, a liquid mass subject to a pressure less than the critical one. Moreover, let us suppose t h a t the liquid is at rest and at saturation temperature. If the temperature of the piston is increased by a few degrees, a vaporization process begins across
120
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
the liquid surface near the piston (see [33], p. 491) and a vapour phase is formed which has a density much less than the liquid one; consequently, the total volume occupied by the system increases proportionally to the quantity of m a t t e r supporting a phase change. Under conditions on phase transition, we can apply the methodology we have followed in previous section with certain necessary modifications. In fact, with respect to the case of solidliquid state change there are two fundamental differences: i) the vapour phase cannot be supposed to be incompressible; it) the densities of the two phases are so different from each other that the vapour density can be totally neglected with respect to that of the liquid. In other words, the two cases differ
for
the
state
equations
describing
the
system
and
for
the
approximations we may introduce. W e again limit ourselves to the case of fields depending on one spatial variable x by
besides the time t. The interface is plane and its motion is given
the function
s(t);
finally
the liquid is incompressible.
Under
these
assumptions our system is again described by the equations (5.17)-(5.19) and it remains to find the further simplifications deriving from ii). From (5.19)j we obtain
P~ P-PlPi .
vv== ~±s, p *.
(5.35)
where the fields without suffix refer to the vapour phase. On the other hand, the vapour density is always negligible with respect to liquid one instance, at atmospheric pressure and temperature of 100°C, the density is 0.958 g / c m
3
whereas the vapour density is 0.596-10 ~
3
(for water
g/cm3).
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
121
Therefore, equation (5.35) can be written as
« = - £ * .
(5.36)
This last relation permits t o eliminate v from the other j u m p conditions (5.19) which become
lvl= -PiPi's\
(5-37)
5 * ( # ) 2 * 2 + p»(M+?)* = - [ * * „ ] .
H £ ) 2 * 2 + M + ? = o. As in the case of melting of a solid body, we can obtain a simplification of the evolution equations by a process of nondimensionalization. In choosing the reference values, the difference between the densities of the two phases is the determining factor. In fact,
it follows from it) that there exists a) a
corresponding significant variation of the linear dimension of the volume occupied by a given quantity of matter when it passes from one phase t o another; b) a noticeable difference between the vapour particle velocities and the interface velocity. These remarks impose the choice of scaling quantities for the length, density and velocity according to the phase. In particular, we scale the densities with respect to their values p ( and p a t atmospheric pressure and vaporization temperature. Owing to the incompressibility of liquid phase, the density will have the same value for x £ [ 0 , s ( i ) ) : p f = p ( .
122
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let
us
introduce
the
nondimensional
P
rate
~Pi
the
water
a = 0.622-10 ~ 3 ) ; the ratio of the linear dimension of the regions occupied by a given mass of m a t t e r in the liquid phase to t h a t in the vapour phase is about \i. Therefore we fix the values Xt and X of the reference values in such a way t h a t Xl coincides with the initial dimension of liquid phase and X =
Xxlii. If we assume that the quantity T — X^/a
, where al = kj/pjCj
diffusibility of liquid, as the time scale, we obtain two velocities Vt = and V = X/T; velocity
is the XJT
we refer the vapour velocity to the former and the interface
to the latter.
Finally, we obtain
the following
nondimensional
variables * — ±, rri , t* =
x x
*»
s(t) Sit)
~ x,
xx
*e[o,«(<)) * e [0,5(0) ;
X -x-s(t) s(t)
r . . , , ., xe[s(t),L{t)],
*= - jt'x,'
+
x
(5.38) (5-38)
+
1T ^T~^ ' *e[«(0.i(0]. P ■ * s * V v =F, P ~J> V,' *
9
P ~ p >
- ee ''
pP*-
s p
pClo'
v
~Vl' e _
C / e'
~v
* ~ c ,e '
where 9V is the vaporization temperature at atmospheric pressure, 0 is the difference between the m a x i m u m and minimum of the boundary temperature and C; is the specific heat of liquid. The choice of the reference pressure is justified by the consideration that the constant pclQ is of the same order as the atmospheric pressure P (for water when we put 0 = 40°C, we have P/pClO~ ~ 1.1). P/pc,Q 1.1).
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
123
Simple calculations lead us to the following system where, for the sake of simplicity, the nondimensional fields have not the asterisk: p= = p(t), p(t), zx€[0,s{t)), e[0,s(<)), 9
H
=
d
>xx
Pt + (pv), ( H .x x = = 0Q,' Pt + Pix~ P'x=
£t e6 [[0,00) 0,oo), '
xe(s(t),L(t)], xe{s(t),L(t)],
-Ap(v, -AP(vrt t
Ve P(eH + ve 'x)= 'x)=
(5.40)
+ vv, x) i, + VV >x)
-P^'x + ^ ' x x
v=-j, » =-*«
'
x = s(t) x= s(t), ,
bl =
(5.39)
«
(5.41)
p 1
1 A^ +(lej ±A°l ±AJ£ (le]++P)s P)s=(9, -/*(*,,) +f(W - ^ ( ^x)+ x ) ++ ,. +0* =(9,x)- x)- -pk(9, 2 =
i 2
(
«
•
.
.
)
■
^4+ w+f= : 0 •
Here we employed the notations of (5.29)2 and A is defined by 4
V2 Cj
The number A is of the order of the ratio between the kinetic energy and heat energy per unit volume and it is very small (for instance, for water, when we put X, = 102 cm., 9 = 40°C, we have /1=6.10 • 10 " 1 3 ) . In view of
124
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
(5.40) 2 and (5.41)2 we see that the pressure field is uniform throughout the whole system, is continuous across the interface and is equal to the external pressure pe. Finally, the system (5.39), (5.41) assumes the approximate form
»„ = » , „ , P,t
xe[0, «(*)),
+ (pv),x = 0,
te[0,oo),
xe(s(t),L(t)},
(5.42) (5.43)
P(e> t + ve, x) = ~ Pv, x + ake> xx i v= - 1 ,
x = s(t) ,
(5.44)
([e]+ £)* = (*, ,)--(<**„)+ ,
W+f = o. This system should be supplemented by the appropriate initial data for the density, velocity and temperature as well as
boundary data for the
temperature and the continuity of this field across the interface. When we assign constitutive equations for liquid and vapour phases we obtain a boundary value problem in the unknowns p(x,t),
v(x,t),
9(x,t) and s(t).
5.4 The case of a perfect gas.
In this section we analyze the system (5.42)-(5.44) under the condition
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
125
that the vapour can be considered as a perfect gas. More precisely, we assume that the nondimensional specific internal energy e and entropy s of liquid are given by the following constitutive equations (^ + 0 ) 9)-e - e +m,, e= = (8 )t +
«, = l n ( 0* .. ++ 0 ) - «, .. ,,
(5.45) (5.45)
where 8^ = 0V/Q and e t , s ± are constant whereas for the vapour phase we have ec = c(fl. + 0 ) ,
S S
(5.46) (5.46)
= cln(9* + 0)-J^]np, 0)-J^]np,
R is the universal constant of gases, M the molar mass and c=c/ct.
The
additive constants have been omitted in (5.46) since it is always possible to assume that they vanish in one of the two phases. We need the constitutive equation of pressure in order to complete the description of vapour:
pP = = ^p{e J% By recalling the definition
* ' .*++•(>)■ )■
(5.47) (5.47)
of specific free energy ip — e - (6^ + 0), the
condition (5.44) 3 assumes the form RE>
(MCA /Mc,\
/Mc-M /Mc-Mc, Cl
+ R\ iiv,
, ,. „v1 e -ln(9^ + 6)}-8^ + ^
= 0.
This last equation can be solved with respect to p to obtain Mcle^
(e^ ++e)--<e~ pp = = CC{$. 0)-fe
R{e R(K ++ e) 6)
*
(5.48) (5.48)
126
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where 7 = - (Mc - Mct - R)/R
and C = Rexp[y +
(Mc,sjR)yMc[.
The inverse function of (5.48) defines the vaporization curve which associates the vaporization temperature to any external value of pressure. T h e density is then given by (5.47) and we can conclude that the temperature and density on the interface are known functions / and h of external pressure pe as follows 0(s(t),t)
= f(Pe),
p(s(t),t)
= h(Pe).
(5.49)
These functions do not depend on the motion of the interface. If we define the nondimensional latent heat by A = ( 0 , + #)[s] we deduce from equation (5.44) 3 the following relation:
A= [e]+£,
(5.50)
which, by taking into account (5.45), (5.46), (5.47), becomes
A=
/Mc-Mc,
+ R\/n
(—w4—r*+(?)+e*-
<5-51)
This relation shows that A is a linear function of temperature on the interface and therefore a function of external pressure pe. Figures 16 and 17 show, respectively, the vaporization curve (5.48) and the latent heat curve (5.51). The characteristic values relative to the vapour-water case have been assigned to the parameters; moreover, the constants e„ and s 4 have been fixed by imposing t h a t the vaporization curve contains the point corresponding to 100°C at one atmosphere where the latent heat assumes the experimental
127
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
value 2.26 • 10 _
10
erg/g.
It is interesting to note that condition (5.44) 3 can be put into the form
g(e,p)-4>,(e) = o where
g
is
the
thermodynamical
Gibbs relations
potential
of
the
(5.52)
vapour.
By
(3.45) 3 and (4.7), we obtain
equation (see (4.13)) dp
p\{9)
de
'* + <
fig-18
recalling
the
the Clapeyron
128
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
llg. 19
We are now ready to make clear the boundary value problem which we are considering to determine the liquid-vapour phase change. We suppose that the initial and boundary temperatures in the liquid phase coincide with the vaporization temperature at a given external pressure; in this hypothesis the temperature field in the liquid admits this constant value on the whole liquid phase. Consequently, we need to solve the problem only in the vapour phase. We start with the remark that (5.46), (5.47) imply Mc,p Mc lPee R
pe = —g—- = pt-— const. const. ,, '
and then , taking into account (5.43)a
2
>> we find we find
CHAPTER 5. STATIONARY AND TIME DEPENDENT PROBLEMS
(pe),
t
+ [(pe [{pe ++ p pe)v> -- ,ik6, nk8, J , x = 0 ..
129
(5.53) (5.53)
This equation can be integrated with respect to the variable x and the resulting arbitrary function depending on time only is determined by (5.44) r In this way we obtain kR i r XR M -vu - (Mc^+ R)pB e,xn(Mc ]i1"+ R)p P = W^P ^W^WP^ -
(5.54) (5 54)
"
By employing (5.47), (5.54) and (5.43)2 we attain the equation eeuH + [Ae,x + [A9,
+Bs}9, Bs}e,xx == A 6A9, , x xxx ,,
+
(5.55) (5.55)
where pkR ~(Mc(Mc + R)pe ''
A
R—
^^
(Mc R)pe (Mc ++ R) Pe
R(8m + 6) Mcp Mc ■ Pee
l(5.56) Dj
°-°
It is quite clear that equations we have not yet used, that is (5.44)2 and (5.55), are not sufficient to determine the temperature field in the domain Q. = fi = {(as, {(x,t):s(t) t):s(t) < <x
(5.57)
because the boundary L(t) itself is an unknown. Therefore we need a further condition which is supplied by the global conservation of mass: 1(0 s(t)++ s(t) S(t) 0 »(
p(x, =J Jht ,, p(x,t)t) dx =
(5.58)
130
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where Jk is the constant mass per unit area and p is expressed by (5.47). Finally, to describe the evaporation of a liquid we have to find the unknowns 6(x,t),s(t)
of the following boundary value problem in the domain
(5.57) 0,t + [A9,s + B8]6,x
x(e)k = n McMr s(t) + -Pe l
i(t) dx
em+&(x,t)
s(t)
=9 ,
(5.59)
(x,t)en,
-fike,x(s(t),t),
R
6(s(t),t)
,
= A0,xx
-Jt,
)
0(L(t),t)
= 01(t)>0
,
where A and B are given by (5.56) and 6 is the nondimensional vaporization temperature implicitly defined by (5.48) in terms of external pressure p .
Chapter 6
PHASE CHANGES IN MIXTURES
6.1 Balance laws i n classical mixtures.
In this chapter we propose to extend all the results obtained in previous chapters to binary mixtures of classical fluids. This further generalization of the model of a continuous system with an interface which we have proposed so far, will permit us to attain the Gibbs rule of phase equilibrium in a mixture, to describe the evaporation of a component of a mixture into a gas, etc. Moreover, it will be employed in the next chapter to supply a model of crystal growth. T o formulate the new model of a continuous system with an interface ( C 1 , C 2 , E ) which we are looking for, we start with modifying the balance equations we derived in chapter 3. In fact, those equations were valid for a system formed by one constituent and now we want to take into account a binary mixture of fluids. If we limit ourselves to classical mixtures (see, for instance, [34, 35, 36, 37, 38]) it is sufficient to add: i) the mass balance of one constituent to the other balance laws; ii) an extra-flux of energy due to the diffusion to the heat flux. More precisely, we suppose that, besides the mass conservation, the following mass balance for one constituent (and then
131
132
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
for the other) is valid:
A -2- ( dt
pv dv+ pi/dv+ Vr
• da) — q-N dapgsvi> q-N da- q -vq^v^dl s sda) s Edl
* Vv Vv n n N- # ) - L dlP v,VVs ■ "E dldl > + », s(.i NN — + Ps ~ C:n„ ■ ) ~TN dl ~ j Ps '"* ' ' *E ' 9
(6.1) (6J )
d9(7
C7
where V is any material volume, a = E n V, v and vs denote the concen trations of one constituent and q, qs are the diffusive flux vectors in V - E and on
£ , respectively, n and JV are the normal unit vectors to E, J> S is the
normal unit vector to da in the tangent plane to E and finally the last two integrals are suggested
by the considerations of section 3.1. T h e
local
equations, which we derive from (6.1) by adopting the usual procedure (see (3.4)) result in
{pv)' + pv V-w {pv)' V - v ++ V V-?g = 0 ,
iinn C - EE, ,
(6.2)
6 2Hc nPs ^ s + St (Ps '.) + V,-(*V, v.)~ -IpvU-q-nj ==
%
T>lnx({C-V)p,v T- n x {{C-V)p svfm-qJ]
= Q, 0, -1s)l =
on T,
where C = Cj U C 2 . By taking into account the local implications (3.24) of mass conservation for the whole mixture, we have
pv+V-q pv + V ■ q==0,0 ,
(6.3)
133
C H A P T E R 6. PHASE CHANGES IN MIXTURES
Psv'+ V,-1.-lP{v-".)U-*■"] T-[nx((C-V)pava-qa)}
onS-T,
= 0, = 0,
onT.
where we have used the derivative operator given by (3.44). With respect to the balance of energy, we recall that in classical theory of mixtures the extra flux of energy due to the diffusion is - p^q , where the coefficient p1 is called the reduced chemical potential . Consequently, it will be sufficient to substitute - h with -(h+p^q)
and - hs with - (hs + Pisqs)
in (3.34), (3.35), (3.36) to obtain the new local equations of energy balance. Moreover, since we deal with binary mixtures of fluids the stress tensors in bulk phases and on the interface should be represented by
t=-pl
+ td,
(6.4a)
T = yl, + Td,
where I = (5-) , Is = (aag) are the identity tensors in the volume and on the interfacial surface E; while p and j denote the pressure and surface tension. Moreover, p, td ,y ,Td are isotropic functions of their arguments, which are listed in the next section. On the basis of these considerations we can write the local energy balance equations of a binary mixture of fluids in the form
pe+pV-v-tr(td®Vv)
+ V-(h+ii1q) = 0 ,
inV-S,
(6.4b)
134
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
paE>.-\p(\{v-V?
yaa^afi
- Tfa{Q0)
+ Va • (A, +
Hsqs)
n
+ e - £ 7 + £ ) ^ - ( A + / i i g ) . n ] + [ P ] ( c n - 7 „ ) = 0,
2
T-lnx[(C-V)pa(±(v-V)
+ E) + T.(V-C)-(hs
onE-T,
nsq8)}}=0,
+
on T. In order t o write (6.4b) 2 we have again used the derivative operator (3.44), the constitutive relations (6.4a), the obvious identities (see (3.50))
«a'(T/s + Trf)T,a =( K / 3 + ^
)
V
)
,
and finally (3.51).
6.2 Constitutive equations and dissipation inequalities.
In the classical theory of fluid mixtures the variables appearing i n the constitutive equations are t h e velocity gradient Vw, t h e density concentration
v, t h e temperature
p, t h e
0 a n d t h e gradients of density a n d
temperature. For the sake of simplicity we assume that they depend only on the first gradients of density, concentration and temperature. Similarly, we suppose t h a t the constitutive equations relative to the fluid mixtures on the interface depend on
and V s #. The
dependence of a is introduced t o take into account the adsorption which might take place on the interface (see section (3.5) and remark i) in section
135
CHAPTER 6. PHASE CHANGES IN MIXTURES
(3.7)) which is now fluid. In order to derive the restrictions on the constitutive equations imposed by the second principle of Thermodynamics [39] we have first to find t h e reduced dissipation inequalities. By eliminating V • h between (6.4) and (3.39) we have -(V> + s$)-pV
■ v + tr(td ®Vv) + V-(ntq)-lh-V0
> 0.
On the other hand, by mass conservation (3.24) we obtain
V-v=
-lp;
moreover, from (6.3)2 it follows that
V • (MX«) = q ■ V/ij + /i a V • q = q ■ Vp1 -
ppxv.
Then the previous inequality becomes
_p^
+ sb-£-p-p1v)
+ tr{td®Vv)-q-Vp,l-\h-V9
P
= 0 .
(6.5)
°
In this inequality ip , s , p , ^ , tj , q and h are functions of Vw , p , v , 9, Vp, V ^ , V0 and consequently
dtp-0, dip . dtp dp dv
'> = dv
♦ & ■ < * • > ■
♦ & • < " > ■
♦ & ■ < * " > ■
♦ & • < * • > ■ ■
136
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
By substituting this expression into (6.5) we obtain an inequality from which it is possible to derive with a standard procedure the following relations
tl> = rl>(p,v,e), l
a\
dip
(6.6) t
a\
2 dip
s = s(p,i/,e) = —jjfg, p = p[p,v,o) = p j - , Pi = t*i(p,v,8) = -fa » tr(td ® Vw) - q ■ V/ij - i h ■ V6» > 0 . We can also extend the procedure which we applied to inequality (3.62) to (6.6)5. T °
t m s en
d'
w e re
gar(l
tne
left hand side of (6.6) 5 as a function /
of Vw, q, V9, provided that we assume an explicit constitutive relation q — q ( V ^ , . . . ) with respect to Vpj. Function / assumes its minimum when all these variables vanish, that is, at equilibrium of bulk phases. Consequently all the partial derivatives of / vanish when they are evaluated at equilibrium:
t°d = (V/Mo = h° = 0 .
(6.7)
It remains to proceed in a similar way with the surface energy balance (6.4) 2 . By eliminating [A] ■ n and V s • hs between (3.39)2 and (6.4)2 , we obtain - p . ( * ' + S9>) + 7 a % + T A
M )
- V • (plsqs) - i h,- V.0
+ [p(j(i'-T02 + s - * - / i 1 > - O ) ^ ] -[pKcn-VJ-I^gJ-n^O,
(6.8)
CHAPTER 6. PHASE CHANGES IN MIXTURES
137
where g is the specific Gibbs potential (3.53). On the other hand, by recalling (6.3)2 we have V V l» -«■ . % - ■VV •. ((MiA) ^Pu l s - ■-^ Hs MIA) = = --9% s "■ P V -v y)U-qU = -9-9 = Hx.P„v' lp(^ - s s) - « ■»], •»]. l *l s ++ A*I,P, " ia- -P (h l sSM S-S ' ^V/i V
and then (6.8) becomes a CT - /»,(«' S0' - ^Hyl sX />.(*' + SO' ) ) + ya ^afi 0a/3 + • Vfl TO<*P ^ 1 lsS -- \1H*fc, •V + TT°/a d (ap) (a/3)'- q?,s • VV/x ) 2 ++ffg_ -tf-* - iiu{y £ l ./,)]# tf + W 5 ((»„ _" FVf + (Hs -nx)q--"] n]
+w|
-^ls( '- 'S)]
+ (Pi S - M l ) «
(6.9)
-[P](c )>0. n-Vn-V n)>Q. -[?](*„ On the other hand according to (3.44 ), (2.21) and (2.14) we derive a' =22aaaa««VVMM))--((cc„„--VVnn))66a a/ /33)) a ' = da=
(6.10)
so that, we write in view of (3.56) a a P's= -VnC))n - ^ n ) ) + /»:= --Ps^ PS% « afi)-Kfi(c V M ) - nM + [pU] I ^ l -.
(6.11) (6.11)
Collecting the results expressed by (6.9), (6.10), (6.11) as well as taking into account our constitutive assumption r af}, V $tf == «(o, p„• v"«»*»« , ^s ,. VV/ )/ ) «(a,/>. s, 6, cr a/»»VsPs PS>s ,vVs^.
we have g = 22 St ** '' = da
aaa/
{Q0r(c n-Vn)baf)) V( a /aa^(a -VJKp) 3 ) - (*n
138
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
+ §- ( " Pfia\afi)
+ PsH{cn - Vn) + [pU])
•s
+
+§^<+§0' s
9^) ^) + dv7s{VsPs)
+
5v^ (v ""' ) + 5 v 7 ( v ^ '
When we substitute this expression into (6.8) we deduce, again in the standard way, the following thermodynamical restrictions on the constitutive equations 9 = 9(a,Pa,va,6),
S = S(a, Ps,v„6)
= - | |
7 =
T
d\afi)
,
(6.12)
/ i l s = nls{a,p„ vt,*) = g^-
2 5* , o
<9*
-PsWS2^^*'
~ Is ■ V s ^ l s - i A. • Vs6> + ( M - 2F 7 )( C „ - V J
+ [p[5("-^)2 + 3 - ^ - / i 1 > - ^ ) ] ^ ] + [ ( / i l s - ^ 1 ) ? ] - n > 0 ,
(6.13)
where we have introduced the surface chemical potential (see (3.59))
".=* + '«^ = =4" (p '* ) -
(6 14)
-
Now, in the usual way, we regard the left hand side of (6.14) as a function / of (T(a/3y qs, V s 0, ( c „ - F J , J7 + , t / ~ , q+,
g~, provided that
we assume an explicit constitutive relation qs = 9 s (V s ^ l s ,...) with respect to
139
CHAPTER 6. PHASE CHANGES IN MIXTURES
V s /x ls . Since / assumes its minimum at thermochemical equilibrium (see [35]), all its partial derivatives vanish when they are evaluated at this case. Consequently, we have [39]
on E at thermochemical equilibrium :
T° = 0 ,
(Vs^f
=0 ,
[I fP JJ(J ] o - 2—7 1t — f = 0 ," +
h°s=0,
?
[9
-fis-^ls(l/+
-^JJO^0 '
[9-
-p,-ih,(v~
-(/s)]o = ° .
(Mi + )o-(^i s )o = 0 '
(6.15) (6.16)
\~-~~/ (6.17)
(6-18)
(fh )o - (//i*)o = ° ■ The meaning of (6.15) t
3
is quite obvious; condition (6.15)2 imposes that the
reduced chemical potential of the interface of the mixture is uniform on E. Owing to (6.18)'s , which imply
[/*1lo = 0 ,
onE,
(6.19)
it is found that the reduced chemical potentials of bulk phases are uniform on E. Finally from (6.17), (6.18) we derive
[ j - ¥ l o = 1'
(6-20)
At non equilibrium we can proceed as at the end of section 3.7, that is,
140
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
we can suppose t h a t the coefficients of v
q
U +,
U~,
, q~ , in (6.13) are functions at least of the second order of these variables.
This implies
that
we can accept
(6.15)-(6.20)
also at
non-equilibrium
provided t h a t the aforesaid variables are not too large.
6.3
Phase equilibrium in a binary fluid mixture.
In this section we want to analyze the equilibrium boundary problem and to verify t h a t the Gibbs rule is satisfied also for nonhomogeneous fields. As we did in the case of phase equilibrium of pure substances, we suppose that the body force admits a potential U, that is / = - VU. On the other hand, dp /dp > 0 and
consequently
the function
p = p(p, v)
is invertible
with
respect to p ; then we can adopt p and p as independent variables. Moreover, from the relation
g(p,v)
= ip(p(p,i/),v)
+
p{v,v) '
we derive
dg__dj)_dp_ dp~dpdp dg _ dip dp dv dp dv
+
dip dv
i _p_dp__\ P p2dp~pp. dp _drp _ pi dv ~ dv ~
l
,fi91, ° lj
CHAPTER 6. PHASE CHANGES IN MIXTURES
141
Finally, it follows t h a t I V p = V - pVv
.
(6.22)
In the absence of surface body force, from (3.27) 1 ; (6.22), (3.27) 2 , (6.7) and (6.15)-(6.20) we derive the equilibrium boundary system
F
i(*> Pi> vi) = 9i(Pi, f ,0 - P-i^i + U{x) - c ■ = 0 , in C- , i = l , 2 Ft(x, p . , i/.) = gt(Pl, i/,.) - / p_ , ^.. +\ £/(*) - nc, = 0 , in C- , i = 1,2 _
[p] = 2 T # ,
[p] = 2 T # ,
7 = const. , on £ 7 = const. , on E
h-wl P\ ++
[a l9
(6.24) (6.24)
=o i
- t*i,(7ivt)
-^-^is(^
(6.23) (6.23)
+
-
- ^" )J] ]o0 = °° .>
pP = pPee, ,
on 3<9C C ,,
(6.25)
j/s = « ,
on <9E ,
(6.26)
where a and c are constant, <9C is that part of <9C on which the external uniform pressure p e is acting and « is an assigned unit vector. We have omitted in all the formulas the temperature 9 which is supposed to be uniform and given. From (6.23), and (6.24) 2 we derive that the reduced chemical potential p. is uniform on the whole region C; consequently / i l s is constant on E and, in view of (6.24)j, the surface concentration v$ is also constant on E. The 11
142
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
unknowns of this system are represented by the fields p t , u-, the constant c^, ai in both phases, the constant 7 and the field vs on £ and by the equation z(x,y) of this surface. Since the number of equations is 10, provided that p e is given, we can conclude that the variance or the degrees of freedom of the phase equilibrium is 3, namely, p e , 6 and, for instance, v. More precisely, let us suppose that the constitutive equations of g and n1 are such that the system
±
±
/ 1 (( p p ± .l .i ^ / M ) == [^ = 0o,, A b1i] l =-(p±,u± ±)
ffo(P 2 admits at least a solution p
+
(6.27)
==19-^1 0, , == o {g-- P i 1/] = >»*)-= = p~,T>+,u~
(corresponding to equilibrium
with a planar interface); then, if the jacobian determinant of functions / i t / 0 at this point is different from zero, we can say that the system (6.27) defines p~ ,u~ +
as functions of p
,f
+
at least for {p
,v
) in a neighbourhood /
+
of (p ,Z7 ) . When we assume that at a point x0 e E the values of pressure and concentration (PT IVZ)
(p*, v'J) e / ,
we can find a corresponding
couple
such that (6.27) 's are satisfied. In this way we find the constants
ci,ai and then the fields of pressures and concentrations in Ca and C2. The equation (6.24)j with the boundary condition (6.26) defines the only surface containing x 0 .
To prove that (6.27)'s are satisfied not only at x0 but
everywhere in E it is sufficient to note that, owing to (6.23) , the differences on the left hand sides of (6.27) are constant in E . Finally, we derive 7 and »/, from (6.24) 4
5
.
143
C H A P T E R 6. PHASE CHANGES IN MIXTURES
6.4
The influence of mass adsorption on surface tension.
Let us consider the phase equilibrium with planar interface in a binary mixture and in the absence of body forces. The equilibrium system reduces to the following one
FfaPfiVi)
= 9i(Pi^i)-Ci
= 0 ,
fhi(Pi>''i)-ai
=
0
in C- , i = 1,2
(6.28)
'
[ p ] = 00 ,, on E [p] iL A * 1 JI
(6.29)
1
h -(M1V} == 0o,, Iff-/*!"] ..+ v [g+ -n [
p= = ppee, ,
All the
10 unknowns
1..
,. \
-I/I/ss)]o )]o = ° on dC dC .. on
pt-, //,-, 7, v , a^, c ; are constant and are
(6.30)
completely
determined by the 10 equations (6.28)-(6.30), that is the variance of phase equilibrium with planar interface is 2 since it is determined, for instance, by 8 and p . Now we want to take 6 and i/j as fundamental variables and omit in the formulas 6 which is held constant. In this way the system (6.28)-(6.30) defines the surface tension 7 as a function of u1. Now it is experimentally verified (see [40, 41]) that can be an increasing or decreasing function of vv
The first circumstance occurs when vs{v\)
<
v
\
144
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
(negative adsorption), while the other one occurs when vs(v-y) > vx (positive adsorption). In order to obtain this result theoretically,
Gibbs assumed t h a t
it was possible to substitute the actual system constituted by bulk phases separated by a narrow interfacial layer with a fictitious one in which the quoted layer is substituted by an internal suitably chosen plane and the bulk phases are assumed to fill homogeneously the remaining part of the same layer. This plane is assumed to be placed in such a way that Gibbs' relative surface excess (see for instance [41, 12]) of one of the constituents of t h e mixture
vanishes.
Under
these assumptions
Gibbs derived
the relative
adsorption equation (see [13], p.25) dy= -Tdpu,
(6.31)
where T is a coefficient, dip $) 2i. = - g ^ = ^ + ( l - « ' > i . and ^{P\s,
P2S)
ls a
(6-32)
function of the densities of constituents of mixture. The
last equality is an obvious consequence of the following identities
Ps = Pis + P2s
The differential djils
'
*(*>1»> P2s) = * ( P » . "») •
in (6.31) is relative to the function /*i s (tO which is
obtained by expressing ps and v in terms of the variable v,. Our aim is to deduce the relation (6.30). T o start with, we regard all the equilibrium quantities as functions of vv omit the index 1 since we refer only to the liquid phase a n d neglect the dependence on a in the constitutive
CHAPTER 6. PHASE CHANGES IN MIXTURES
145
surface equations. Moreover, we remark that if we introduce the volume quantity d(pil>) J/ Pl=-%r + (l-v)l* " 1 - dp, = 9=9+ (l- - )1^ l ,, we derive from (6.29) 2
(6.33)
3 4
(6.34) (6-34)
h£i = = Khs>. and we can write (6.31) as follows
and we can write (6.31) as follows dj = - r
dp,.
(6.35) (6.35)
d1=-Tdpl. On the other hand, we obtain from (6.24) 5 , (4.54), (6.12) 4
On the other hand, we obtain from (6.24)5, (4.54), (6.12)4 7 ^ P . ^ - f f + PiC*'-".))•
(6-36)
As all the quantities in both members of (6.36) are to be regarded as functions of v, we can calculate their total derivatives as -".)) £ = ^ - ' + "i("-0) £=^w-«+M"dv x „ ((99dV ddP d$ du s* dg dg dp Ps,dV dp s + + K( dl/ \ dPa sdv dv dpdv f dv *{Wp~ ^ dv d^^~d^~dp~
+ + pPs
dp, , u - dv (Ms -Vu s
./d^
dp
d^x
/
dus\\
S ■'J$*+%)+4-T® WT W + ftr¥/i
7 v\ (( g + fi1is + 7 ) + Ps (Ps
[dip dp\ (9±_j fdrp L p i l^\dp dP\ dp ( (djP iidp) -{fo + P dvJ'Kdp f? + Pdp)dv + ^~
°^« 77 dp* 'PI"s
dv ^s
dv
.dj^
(
dj^\\
-J&W'-S))-
146
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Owing to (6.12) 4 and (6.24)5 , we finally obtain dj d-1
d
1i
Pl v» dP-i
Pas dp dp P
= -r-P.\v.~v) rsy dv dv ' -j dv
7r-r-P dv
IR Q 7 N v(6.37)
'
In order to compare (6.37) and (6.35) we observe that (6.33) implies d
px -+(1_j/) dJh
dji H 1=dlddg dg R dp + dg__ d
dv rfi/
3/? dp dv dz/
~\dp~
<9JV dv
' dv df
V , 1 dP \dp * P dp) dv"
'^~dT V ■8)£+®l+*!S)-*+p■ /i a + ( i -
d^ ' /I P dv; ++Kl( l -■^-arThis relation allows us to put (6.37) into the form 1 - - ^ 5 dp
-v dp,
d7 d
U Ps P l 7r Ps=1Ps -f^¥-£- 1/-i^¥- -v dv P 1dv
dv
dv
rs
1 -v dv
P 1 - v dv
y
But in ideally diluted mixtures it is (see [41], p. 333) p ( l -- i /i/)) ddp . (i <€ p rfPl
K-
(6-38)
(6.38)
" 1 s
and moreover (see [53], eq.(96.7), p.319) d£
fdi/ la>>00 ;; dv consequently the function y(v) has the experimentally observed behaviour.
'
147
CHAPTER 6. PHASE CHANGES IN MIXTURES
6.5 T h e Gibbs principle for phase equilibrium in fluid mixtures.
In this section we extend the results of section 4.3 to the equilibrium of a liquid binary mixture C^ in the presence of a mixture of its vapour Cy. We start again from the Gibbs principle according to which the equilibrium configurations are extrema of the total free energy with respect to variations at constant mass. Let
da
(6.39)
be the total free energy of the system where ifiF = ^f{pp,i^p)
, 4>y =
*=
Pvi^v + U) dv + p^s
pF(ipF + U) dv + C^r
Cy
il>y{py,Vp) and \PS = ^s{ps,a,v
s)
and E is a regular surface. Let us introduce
the set of functions K = {k : E —► Jfc(E) ,regular, and conserving the surface mass }, R
= {Ps> vs'-
E
—* *
+
> re g u l ar }»
7V = {z/: C — » » + } , where C = C^ U Cy. Moreover, we will denote by ft : C —► C, a mapping such that: i) it is regular in C - £ and exhibits finite discontinuities together with its derivatives across E; ii) h,p
: C^i —► h(Cp), h,~
: —► /»(Cy), are diffeomorphisms and
locally conserve the mass. If we denote by H the set of these mappings, (6.39) can be regarded as a functional defined over the set H x K x Rx N which can be normed in the
148
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
usual way. The Gibbs principle states that {he,
ve,
ke, pea, ves) is an equilibrium
configuration if *(A e , v€, ke, pes , v\ ) is a conditional extremum with respect to all the arguments h, v, k, ps , vf
$ =
pL dv +
pLvL
$ 1
dv+
satisfying the global constraints
py dv +
PyVy dv +
(6.40)
Au
Ps
p „v „ da IS
S
As is well known, the extrema of (6.39) under the conditions (6.40) are defined by the condition
(6.41)
d(^r + A$ + A 1 $ 1 ) = 0 ,
where A and Xx are numerical Lagrangian multipliers. By proceeding as we did in section 4.3 and choosing the unit vector n normal to E directed from C^ to Cv , it is easy to verify t h a t (6.41) can be written as follows
( V p + pVU)
p(p1 + Aa) bv dv
-6hdv+ C
p + (g
p~(g
+
U + \ + \1)
(6hn-6kn)+
da
+ U + X + X1)~
{Shn-6kn)~
da
+
(6.42)
CHAPTER 6. PHASE CHANGES IN MIXTURES
149
+ 17, a &kada + | ([p] - 2Hj) 6kn da + (Vs + Xl)
6
"sd
E
^ ( p A +A+ A l ) V s d ( T = °
+
where p =
p
Jp-' „_0V
^-^7' / , P
9 = ^P + j ,
2 SI'
7 =
^,
It is now quite evident that we again
9
„ „ 3*
(6.43)
„ _ d#
^ - 5IT ' ,T,
3*
find
fl(p.*) " />s "
the equilibrium system (6.28)-
(6.29).
6.6 The evaporation of a fluid into a gas.
T h e previous considerations are relative to phase equilibrium in a binary fluid mixture. In this section we want to study the evaporation of a fluid into a gas [42]. T h e system with an interface we use is made up of a pure liquid phase Cj and of a gaseous phase C2 which, in turn, is a mixture of the vapour of the substance filling Cj and of another gas (for instance the evaporation of water in the presence of air). The dynamics of this process is carried out by supposing t h a t the following hypotheses are verified: a) the phenomenon
150
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
exhibits plane symmetry; b) the pressure is less than the critical one; c) the interface
has
no
material
characteristics.
These
hypotheses
imply
the
following approximations: i) the fields depend on the temporal coordinate t and on one spatial variable x; ii) the liquid phase is incompressible and at rest; Hi) the mixture is made up of perfect gases; iv) the mixture density is negligible with respect to that of pure phase. In these conditions the region Cj is represented by the interval [0,
s(t))
and the region C2 by the interval (s(2),/(£)], where s(t) is the abscissa of the planar interface and l(t) denotes the abscissa of the free planar boundary of vapour mixture. Moreover, the balance equations of mass (3.24), (6.3), of m o m e n t u m (3.26) and of energy (6.4) assume the following form
w = 0,
P,x = 0,
* e[0,«(<)),
0, t -
a
i
,XX
<e[0,oo),
(6.44)
'
where aj is the liquid thermal diffusibility (see (5.22)) and
P,t + M . * = 0 .
* 6 («(*),/(<)],
*e[0,oo),
p(uH + w,x)=
~q,x
,
p{v,t + vv,x)=
-p,x
,
p(e, t + ve, x) = - pv,
+ kd, xx - (// 1 ? ),
,
(6.45)
151
CHAPTER 6. PHASE CHANGES IN MIXTURES
where k is the thermal conductivity of vapour and
p(s-v)
= pls,
x = s(t) ,
pv(s-v)-pls-q
<e[0,oo),
(6.46)
=Q ,
pv(s - v ) - [ p ] = 0 , 2
P(^v
+ e + ^) (s -v)-
Pl(e+^)s
1 v2 + [V> + ?]"-pxv2
+{ke,J-
p,q = 0 ,
=0 .
In writing (6.46) we have recalled that in the liquid phase v = 0, v = 1 and g = 0. Since p j ( see section 5.3), from (6.46)1 we have
r ~
.; s ,
(6.47)
P
and (6.46)2_5 become g = p/*(i/-l),
(6.48)
ip]=-4*a. />/«
/IP?
\2 „2
s2 + W-
►?)*■[**„]--HlPls(u. -1) =, 0 , !
! ( # *
+M+J--Hv-
=0.
Now we can repeat the nondimensional analysis of section 5.3 without any modification and attain the following system
152
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
v = 0,
p,s = 0,
* e [ 0 , *(*)),
eH = />.t + ( H - x = ° .
te[0,oo),
(6.49)
°1 XX '
*e(«(*)i'(*)].
p{y,t + vvtx) =
<e[0,oo),
(6.50)
-9>i .
P»x = 0 . P( e . t + »e> x) = - P v ' x + a ^ ' xx - (M),
v= - i ,
* = «(*).
X
,
*e[0,oo) ,
(6.51)
[p] = 0 ,
(w+J+a-"))* = (»,«)"-(«w,x)+, T V* -")}°
—
\"ixJ
— 1 U«/l/ ,
M+J+("-i)/h = o , where a is the ratio of the vapour density at atmospheric pressure and vaporization temperature to p ( ; moreover, we have used the same symbols for nondimensional quantities. To the previous equations we have to add the initial data and boundary conditions:
s(0) = s0 , v(x, 0) = w0 ,
P(x,
1(0) = lQ ,
0) = p 0 , v(x, 0) = v0 ,
0(o,t) = * i ( 0 . p(l(t),t = pe,
9(x, 0) = 0O ,
o(i(t),t) =
v(l(t),t) = l ,
(6.52) xe [0,1(0)} ,
e2(t),
q(l(t),t) = 0.
pe being the external uniform pressure. The boundary value problem (6.49)-(6.52) in the unknowns p(x,t),
i/(x,t),
0(x,t),
v(x,t),
s(t) and l(t) is very complicated; however, it is still
CHAPTER 6. PHASE CHANGES IN MIXTURES
153
possible to derive interesting results from it. First of all, we recognize at once that p(x, t) = pe for any x and t. On the other hand, when we look at hypotheses
«), Hi), the free energies of liquid and vapour phases are
respectively expressed by the following constitutive equations
i> ^(8) 8)[l -- ln(0* ln(0*++8)) 8))-- ee00 ++ (8* (8* + 8)s00 ,, t(8) = (8* + 8)[1
(6.53) (6.53)
V>(0, P,v) p, u) == [cu + c (1-i (l- »w i/)](0*+*)[! + 8)[1-ln(0* ln(0*++0)} 8)} m A
+^ ^v{8* + 8)\npu ^npV + ^ R- (- 0 l -- ■u v){9*. W ¥8)\np{\ + 8)lnp(l-u), + -»), + c,M 4 R
where R is the gas universal constant, M and M A the molecular masses of the evaporating substance and of the mixture respectively, c = c/ct and cA — cA/ct
are the ratios of specific heat of vapour and evaporating substance
and liquid; finally, 9* has already been introduced in section 5.4. By the thermo-dynamical relations , . e = if> + 8s , p p
-w
d\p 2dip
dtp dip s — 88 '
^
=
dtp dip
dU'
and (6.53) we can put (6.51) 3 into the form
P , ==/ where A,j
M
' A'
M 1 *"+ g M - "l -"\e( W+»)-<,->'+» ) - ^
(6.54)
and 6 are constants. The external pressure being fixed, Eq. (6.54)
supplies a relation v = v(8) between the temperature 8 and the concentration v on the interface. In particular, at ordinary pressure (p = 1), the evapor-
154
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
ation temperature corresponding to a given concentration v is obtained by intersecting the curve 7 „ y = A{6* + (?) ~ 7 e
e* + (
(6.55)
with the straight lines MAv
MAv + M{\-v)
'
(6.56)
In fig.20 the curve (6.55) is represented together with the straight lines (6.56) for different values of v under the assumption that the physical
Fig. 20 parameters refer to the water evaporation in the air. The figure is, in agreement with experiments such that the evaporation temperature decreases
155
CHAPTER 6. PHASE CHANGES IN MIXTURES
with the vapour concentration on the interface. If we define the (nondimensional) latent heat
A = <0*+ * ) [ « ] , condition (6.51) 3 becomes
(6.57)
{e] + % + (\-i>)fi1=\,
and (6.51) 2 can be written as
\s=(9,x)-
-(ak9,x)+
.
This relation is formally similar to Stefan's condition that we have already deduced in the case of melting (see (5.34)). However, A in (6.57) is not constant but depends on the unknown temperature on the interface. Now we want to show that the boundary value problem (6.49)-(6.52) can be further simplified if the gases in the mixture are both monoatomic or diatomic. T o verify that, we begin by writing the balance of energy (6.50) 4 in the vapour phase in the following form
(pe), t + [(pe + p > ] , x = (akO - / i l 9 ) , x .
But p and e are expressed by the relations
(6.58)
156
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
_ MMAc, p ~ R [MAv + M(l-v)](e*
p
+ 6)>
l
°-°9j
e = [cv + cA{l-i>)]{8* + 8), which can be easily derived from (6.53). Consequently, if the gases in the mixture are both monoatomic or diatomic, the specific heats c and cA inversely proportional to the molecular masses M and MA
are
and then the
energy E = pe and the enthalpy H — pe + pe per unit volume are constant and uniform. From these remarks and Eq. (6.58) we derive
v=
ak9,x-p1q H
+f
(6.60)
where / is a function of time which we must determine starting from the interfacial data. Relations (6.59) and (6.60) can be introduced into the mass conservation (6.50)j and the partial mass balance (6.50)2; simple calculations lead us to the following system of nonlinear parabolic equations in the unknowns 9 and v
E\MAv
+ M (1 - v)}6, t + H(MA - M){9* + 9)v, t
+ (ak9, x-piq
+ f)[(MAu
+ M{\ - v){9* + 8)},x
= [MAv + M{\ - v)}{8* + 0){ak9, x -
%g)
,x (6.61)
HvH + {ak9,x-q HR[MAv
+ f)v,x
+ M(l-v)}(9* MMActpe
+ 9) ~q,x
'
157
CHAPTER 6. PHASE CHANGES IN MIXTURES
Finally, we need a further equation in order to determine the free boundary l(t); such an equation is supplied by the global mass balance that can be written (see (5.58)) as l(t) s(t)+
p(i>(x,t),0(x,t))
dx = const ,
(6.62)
s(t)
where p(is, 6) is expressed by (6.59)1. To
summarize
we
can
state
that
the
temperature
9(x,t),
the
concentration v{x, t), the interface abscissa s(t) and the free boundary position l(t) represent the unknowns of the following system
i£[0,s(()),
te[0,oo), 9, t =
xe(s(t),l(t)}, re(*(0,/(0].
o, XX
(6.63)
'
t* e6 [[0,oo) 0 , o o ) ,, r r r ?IT ,. . s ^ / 1 ,.\1a , TJI H[M Av + M(l - v)]6, t + H(MA - M){6* + B)v, t
+ (ak9,x-li1q
+ f)[{MAv + M(l-v)(B*
(6.64)
+ e)],!t
= \MAv + M{1 - v)](8* + 6)(ak6, x - Hq) , x ,
H»,t x = ${t) ,
+
, ,„ i«ke,x-q
+
,, f>,x=
HR[MAv + M(l-i/)](6* MMAclPe
+ e) "«». •
t e [0,oo) , \s=(e, ake,x)+ ,, x)+ A* = (x)--(akO, * „ ) - -(
(6.65)
158
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
with boundary and initial conditions
q(s(t),t)
= Pls(v-l),
q(l(t),t)
0(0,0 = 01,
= 0,
(6.66)
*(J(t).0 = W .
= Wh- *
/("(«(<). t),9(s(t),t))-; tf(ar,0) = ff0(r) ,
+
-J/)^
= o,
=
v ( i , 0 ) = i/„(x) ,
«(0) = s0 ,
/(0) = / 0 ,
and the integral condition l(t) s(t)+
p(v(x,t),0(x,t))
dx = const.
(6.67)
s(t)
In (6.66) 5 ip is given by (6.53) and p in (6.66) 5 , (6.67) is obtained by inverting the constitutive equation p = p(p,v,9)
= pe which, in turn, is given
by the thermodynamic relation
P
2dip
= pJp
Finally, we have to give the constitutive relation q(9,v,9,
,i>, ).
Chapter 7
CRYSTAL GROWTH
7.1
Gibbs' and WulfPs equilibrium laws.
In this chapter we are faced with the difficult problem to describe the equilibrium and the growth of a crystal in its melt or in a mixture by resorting to the model of a continuous system with an interface. We shall see t h a t this approach has to be modified if we want to derive from the model the experimental evidences. W e begin with deriving the classical results
about equilibrium due to
Gibbs and Wulff [12, 43] although the proofs we propose to obtain them are much more formal [28]. Let C be a polyhedric crystal having a volume V, n faces
and m edges /
(p = l , . . . , m ) . We denote by /■■ the
edge between the faces cri and a ■ and by i/- the unit vector which is normal to /•
and lies in the face a{. Finally, a uniform surface energy Ei per unit
area is associated to any face ai of C. Moreover, E^ is supposed to depend only on the orientation of c^ with respect to the crystal cell. Let C be a polyhedron and [C] the class of all polyhedrons having the same volume V, the same number of faces and edges of C and faces which are
159
160
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
parallel to those of C. The Gibbs principle postulates that a polyhedron C represents an equilibrium configuration for the crystal if total surface energy
**=
71 n
i1l
£,. **
(7.1)
has an extremum at C in the set [C]; moreover, this configuration is stable if it corresponds to a minimum in [C]. To determine the conditions which characterize the polyhedron C, we need to give the procedure to pass from C to any polyhedron in the class [C]. To this end, we introduce the normal displacements 8hni (i = 1,...,n) along the unit normal n, to the faces ai as well as the displacements 6K-
of any edge I-
of C. Of course, these
displacements will define a new closed polyhedron in [C] if the following conditions are satisfied (see fig. 21)
Fig. 21
161
CHAPTER 7. CRYSTAL GROWTH
6KtJ.ni
= 8kni,
6KXJ.nj
(7.2)
= 8kn3
It is quite evident that the class [C] is obtained by varying arbitrarily 6kni and 6K— provided that these displacements verify conditions (7.2) and leave the volume V constant. According to the Gibbs principle we can say that C is an equilibrium configuration for the crystal if the total surface energy (7.3)
*(ttm.Stf,-j)=E,-^*i has an extremum at (0,0) for any (Sk^SK-)
satisfying (7.2) and the global
condition
*(«*„,, tf*y) =
(7.4)
dv - V = 0 , C
V C e [C]. This is equivalent to say that (0,0) is an extremum for \P - A$, where A is a Lagrangian multiplier; that is, (7.5)
d(tf-A$) = 0 , for any (6k itSK:A
verifying (7.2). On the other hand, by Gauss' theorem we
have dv — i (
rf$ = \ dvC
C
aC
(r + 5knn) ■ n da dC
r ■ n dv) = ^6kn dv , aC
162
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where r is the position vector of any point on the crystal surface. By recalling that 6kn is uniform on any crystal face o^-, we can write
d*=±±iHni'i-
(™)
Owing to this last result, (7.5) is transformed to the form m
n
E PiEivij + V i < ) • 6Kij ^ - x E i 6km °i = o .
(7-7)
for any (6kni, SKiA verifying (7.2). In (7.7) A = A/3 and the indices i and j depend on the edge index p since they are relative to the faces ai and awhich form the p-th edge. If we put cosa • =nt-• n , it is an easy exercise to verify that v
ij
~ ~ cotaij
n
i +
csca
n
ij
j
■
(7-8)
By taking into account (7.2), (7.8), we can write (7.7) as follows m
n
E piAJK,
+ AJt6knj) ltJ - A E i *t Skni = 0 ,
(7.9)
where A
ij
=
E
j
CSCa
ij
- Ei
C0ta
ij
■
(7-10)
The first summation in Eq. (7.9) contains the same number of terms as pairs of adjacent faces and it t O CX11U l b becomes l_»C»-.tJlllCD n
1r■ (V
A i
- A
163
CHAPTER 7. CRYSTAL GROWTH
where Yl 'j indicates that the summation is extended only on edges which form the boundary of the face a^ From the arbitrariness of Skni previous relation
we finally
derive the Gibbs rule for the
in the
equilibrium
configuration of a crystal ([12, formula 666])
£-iE'jAijlij
To
recognize
which kind
= \,
« = l,...,n.
of conditions
Eq.
(7.11)
(7.11) imposes on t h e
equilibrium form of the crystal, we analyze with greater attention to the case of a polyhedron such that there exists an internal point r 0 the perpendicular from which t o a face intersects it in an internal point of this face. If the distances from r 0 to adjacent faces cri and a ■ are hi and see from fig. 22 t h a t
fig. 22
h- we immediately
164
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
hh n fH
a
a v
-V>i = o.-
jihji3nJ i i + aijvij-aij''ijjiVji-
=0
Thus we have a ij
a
ji y + °j*
sina -, = h*j i sma ij = ij tJ>
COSQ COSQ-
cosa-j sina.j= h^i sina ji + afly ij cosa ij = ij <, i« +
a a
or cota^■ csca = ji csca ij (J -~--h ^i cota ij >>o, 0, h csca a ta- > 0 . csca aa ji = ij tJ -~- -hhji coc^o ij > 0 ■ H = Ki
a
ijij
a
= hh
7 ((7.12) -12)
We then deduce that /i1 and hJ• should satisfy the conditions ^hj «way < j - ,
cosa.j < j ^ ,
h h ii cosatJ < Y , cosatJ < -^3 ,
which have to be met if two adjacent faces are formed. From (7.12) it follows that °i = \ Z'i « y hj = k E 'i (^ csc a ! J -fcf.cota,-,-) I.. .
(7.13)
Similarly, it can be proved [44] that /■ ■ is a linear combination of h^ h-, h.-_!, ^y + i- Introducing these quoted relations and (7.13) into (7.11) we get a system of n equations in the unknowns ft- which in principle supplies the equilibrium configuration of a crystal for which we have given a priori the number n of faces and the angles a ■ ■ between their normals. The value of A is determined by inserting the values of h^ which depend on A, into (7.4). Moreover, if we substitute (7.13) into (7.11) we obtain
"I*.
Ot 2£ ij [(^[ (E3 -1 cscay ■-- ((££, . ,. - | />, )) Ccota^.] = 0, « , j ] '/,, ij = 2 h^ Vcsca,j
165
CHAPTER 7. CRYSTAL GROWTH
so we conclude that for any crystal there exist possible equilibrium configurations in accordance with the Wulff law 2E- ^ = A,
i = l,...,n.
(7.14)
We can easily find the constant A by introducing (7.14) into (7.3) so that we have *
_
A n 2 i
K °\ ~ 23
AV,
where V is the crystal volume, and thus A = 2\t/3V and has dimensions of a pressure. We can also obtain the equilibrium configurations of a crystal by mi nimizing the total free energy at constant total mass. For the sake of simplicity, we suppose that the crystal is completely surrounded by its melt or vapour and the pressure is everywhere uniform. Then the total free energy can be written as (see (4.43))
* = Ei i £,•*< + pi>{F) dv +
pF^F{pF)
dv ,
(7.15)
C-F
where C^ is the region occupied by the melt or vapor. It is an easy exercise to verify, by resorting to sections 4.7 and to the calculations relative to (7.1), that an equilibrium configuration of the whole system is characterized by the following conditions:
p = const
in C ,
pF = pe ,
in CF ,
(7-16)
166
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
i> + p = w-iE'i^ijUj
4>F
+ PF -
= p-PF>
«' = !»•..,"•
In particular, the Wulff law becomes IE-j^-P-PF^
i=l,...,n.
(7.17)
We conclude with the following remarks: i) In the formulas one finds in classical papers or textbooks there is no difference between surface tension j : of the crystal surface a^ and its surface energy E±. This is due to the hypothesis that ai is like a soap film (see section 4.7, remark i)). ii) Owing to the very small values of the surface energy Ej, we have an appreciable pressure jumps on the crystal faces only for very small crystals (germs).
7.2
About WulfPs construction.
To apply Wulff law we have to give the number n of faces of the crystal. This means that we can have many equilibrium configurations for a fixed jump pressure depending of n. Wulff suggested that there is only one
167
CHAPTER 7. CRYSTAL GROWTH
equilibrium configuration which is in accordance with (7.17) and in addition represents an absolute minimum for the surface energy. This configuration should be obtained with the following Wulff construction. Let us consider the function E(n), which gives the surface energy per unit area relative to a crystal face whose unit normal is n. If we put /(n) = 2E{n)/(pc
- pL), the
equilibrium configuration of the crystal is represented by the spatial region W which is defined as follows W = {re3?3:r-n (n),VnGS2} ,
(7.18)
where S denotes the unit sphere (see fig.23).
fig.23 The previous considerations are not quite satisfactory for several reasons. First of all the Gibbs formula characterizes an infinity of equilibrium confi gurations which could be stable or not and in this latter case they would not
168
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
be observable. Moreover, an equilibrium configuration C is an extremum for the total free energy only in the class [C] of suitable polyhedrons and not with respect to arbitrary variations of size and form of crystal C. Finally, Wulff construction of equilibrium configuration is not justified by proving that it corresponds to a global minimum for the total surface free energy. A correct formulation of crystal equilibrium consists in finding the closed surface E which represents a minimum for the functional *(E)
(7.19)
E{n)da,
under the constraint that the volume V internal to E is assigned and in the hypothesis that the surface energy per unit area E(n) : S2 —► R
+
is a
continuous and a.e. differentiate function (see fig. 23, and [45], [46], [47]). This variational problem is discussed, for instance, in [48], [49]. We remark that this approach to crystal equilibrium attributes to surface tension a dominant role in crystallization but does not permit to relate this process to temperature and pressure. A more complete variational formulation of equilibrium is obtained by requiring that the displacement field « in the elastic crystal C, its boundary E and the density p in the fluid phase £p represent a minimum for the functional
* ( « , £ , p)
ipc(Vu)dv+ C
I if>F(p)+ CF
E{n)da,
(7.20)
Z
under the constraint that the total mass is constant. It is evident that for macroscopic crystals we can neglect the surface integral since the total surface
CHAPTER 7. CRYSTAL GROWTH
169
energy is very small with respect to the energies associated with the volume phases. Moreover, in the case of a crystal which is not in the presence of its melt or vapor but is itself composed of phases, we have to consider only the first integral (see [53, 54]). We want to conclude this section by noting that these last variational formulations do not describe the equilibrium of a crystal in its melt or vapor since they foresee that the pressure jump becomes zero when the crystal faces are planar. This result does not agree with Gibbs' or Wulff rules which in turn justify the supercooling or the superheating of small crystals. In the next section we propose an approach which is based on a new interpretation of nonlocal thermomechanical theory.
7.3 An introduction to nonlocal thermomechanical theory.
The basic assumptions of standard continuous thermomechanics can be summarized as follows i) the general balance laws are valid for any seabed V of the body C, however small it is ( localization principle); it) the physical response of a material particle X e C depends on the state of C at X; this means that only interactions having a very short range are effective;
170
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Hi) the boundary values can be prescribed on any part of the boundary dC independently of the values one assigns on other parts of <9C. We shall call a field theory nonlocal if some or all of these assumptions are relaxed. We do not want to discuss here the motivations which suggest us to leave one of the previous hypotheses; for a deep and exhaustive analysis of the foundations of nonlocal thermomechanics we refer to the quoted references [54, 55, 56]. We are only interested in putting in evidence the particular viewpoint we adopt in this essay which is essentially concerned with the assumption i); to this purpose we begin with discussing some examples. Let us suppose that any particle of a continuous system C interacts with the others in an arbitrarily large neighborhood of it and let / be the relative specific body force. Moreover, let us suppose that this interaction satisfies an action-reaction principle or, more generally, the condition / dv = 0 .
(7.21)
C It is quite clear that / does not appear in the integral momentum balance for the whole system; however, its contribution is present in this balance law when it is referred to any part V of C, V ^ C, or when it is written in local form. This conclusion is trivial but it may be read in a more interesting way. Let us suppose that we do not know in details the physics of the system C so that we ignore the presence of the internal interaction we are speaking of. If we suppose that the momentum balance is valid only for the whole
CHAPTER 7. CRYSTAL GROWTH
system
171
and not for any part of it, we reduce the damage of our ignorance
since a body force / , which satisfies (7.21), appears again when we localize the momentum balance. We shall call / a localization residuals. Let C be a rigid conductor of heat which is initially at a nonuniform temperature 9 and is bounded by adiabatic walls so that it does not exchange heat with the external world. If we do not know that C is able to conduct heat by previous experiences and apply the balance of energy to any part of it we arrive at the wrong conclusion that the temperature remains uniform at any point of C. However, if we suppose that the balance of energy is valid only for the whole system 4d dt
e(6) dv = 0 ,
(7.22) 22) (7
(C 3
we locally have, e= f ,
where
r dv = 0 , C 3
where f is a localization residuals, so that the temperature is not uniform. If, in addition, we suppose that r = - V • h, we obtain again the right equation. This example shows that our loss of information about the internal nature of system C is only partially compensated by the hypothesis that the energy balance law is valid for the whole system and not for its parts. In fact, we have to assign a constitutive equation for f and it is difficult to imagine a priori, for instance, that f — - V • h — V • (k- V#), where k is the thermal conductivity.
172
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let us go back to the considerations we have developed in section 3.2 to justify the model of a continuous system with an interface (Cj, C2, £). In that section we considered a general balance law for some fields / , (p, r which were defined over the whole system but undergo continuous and fast variations across a thin layer A. Suitable averages on this layer led to fields associated with a surface T, in the layer. Let us now consider the case in which ip and r can be neglected in all the region C except for an internal narrow layer A in which they are prevalent. The same average process leads again to surface quantities $, R on E but now these fields do not appear in the global balance law since they verify the condition
$ ■ vv dl + Rdcr = 0
and they will be called again localization residuals. Once more, if we ignore the presence of these internal interactions we can mitigate the consequences of this loss of information by the introduction of localization residuals. It is quite evident that the introduction of such localization residuals as well as their constitutive equations can be justified only a posteriori by the results they permit to derive. This is what we have in our mind when we try to apply the nonlocal thermomechanics to crystal equilibrium and growth. This approach, provided that it leads to concrete and physical results, represents the simplest way to describe these phenomena without resorting to complicate models which, on the other hand, supply a deeper understanding of these processes.
CHAPTER 7. CRYSTAL GROWTH
173
7.4 Nonlocal balance laws.
Let (Cj,C 2 ,S) be a continuous nonpolar system with an interface and let C = Cj U C2; we suppose that the general balance law (3.2) is valid only for the whole system [55]
da) == N-
(7.23) (7.23)
When we localize this equation we have, instead of (3.4), the following equations
f + fVv-V
in in C C --EE, ,
(7.24) (7.24)
+ n -=y>]J R, = R on , Eon- E T -, T , + V ss -■( (F F ® V V) ) - 2- #2Hc c n nFF- V- ,V-s* ■- [$/ t- / [fU + n-v>] -$,, fr-[nx((C-V)F+*)]= [ n x ( ( C - F ) F + f ) J = -§,,
on T, onT,
where the localization residuals f , R , 1R> verify the integral condition £d<7 fi dcr++ % dd// = 0 ..
r? ddv u ++ C
1£:
(7.25)
ir
In the sequel we suppose that one or both of the phases Cx, C2 of the phases we are considering are occupied by binary mixtures. By applying (7.23)(7.25) to mass conservation, instead of (3.24) we derive
174
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
p + pV-v
= p,
+ VVss-(p .(psV)-2Hc -lpU} p''ss + -lpUj nPnP P sV)-2Hc
inC-E,
(7.26)
== p sps,, on on E E - TT ,,
r.[nx(C-V>J=-?,
onT,
where ( )' is defined by (3.44) and the localization residuals p, ps, r satisfy the global condition
p dv + \ps da +
r dl = 0 .
(7.27)
E
Equations. (7.26), (7.27) clearly imply that although the mass of the system is conserved globally there are local creations or annihilations of masses which are compensated by annihilations or creations elsewhere. This situation is not comparable to the one which is encountered in mixture theories where local mass transfers are allowed between species occupying the same point. To our knowledge there exists no physical phenomena within the classical framework which may lead to such mass residuals. So they should be eliminated. Moreover, if we keep them we would obtain in the other local balance equations some non objective terms (see [57]). For the same reasons the mass balance for one constituent assumes the ordinary form (6.3)
pv + V-q = 0,
inC-E,
V i +v , - « . - K " - , / . ) y - « - n ] = 0 , r-lnx(C-V)psV-qJ
= G,
(7.28) on E - T , onT.
When we apply (7.24)-(7.25) to the balance of momentum, instead of (3.27),
CHAPTER 7. CRYSTAL GROWTH
175
we have pv Pv = V-t + pf + pf ,
iinnCC--EE,,
(7.29) (7.29)
p.v PsV' = VVss-T -T r+[ + + lp(v-V)U lp(v-V)U + + n-tj n-t] + + ppssFF + + pr* pssF, F, on on EE -- TT ,, P(v- - V)U + n ■H + P,f+P. T-lnx(C-V) V Psv + + T} T]== -- ## ,, co n T , [»x(<7-Ps-v) where, as usual,
pps5FF doda++
/>/ d« cfo + /?? C
E
<J $ d/ = 0 .
(7.30)
r
For the balance of angular momentum we derive (see (3.30) too): ee1l -<xe-txei + + rrxx ppf/ == p? pi ,, iinn C C -- EE, ,
(7.31)
aa aa ■ • TT xx aaaa ++rr xxppssFF == ppssLL ,, on on EE-- TT ,, - r x #? = -- I£, ,
on onT,
and the global condition
pi pi dv dv++ C e
c* r id/ = = o0 pssLda+ X do- + Ldl : i E r 5
should be satisfied. We remark that rxpf
and rxpsF
(7.33) (7.33) are not objective so
that we can introduce the objective quantities
pi* = pi -rxpf, pi* pi-rxpf
,
PsL* psL* = pPssL-rxp L-rxp33FF
which permit to write (7.31) and (7.32) as follows
,
(7.33)
176
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
= pl* ,
e'-txe{ aa-Txaa
= p.L* ,
(7.34)
inC-E, onE-T,
c p(l* + rxf)
dv +
c ps(L* + rxF)
da +
c s
L dl = 0
(7.35)
r
The stress tensor i £ C - E and T on E are represented by t = f \ ®e■ ,
T = Taf3aa a„ + T a a a ® n ,
(7.36)
due to the constraint n - T — O on E. Introducing (7.36) into (7.34) we see that Hihii3*H + Pi* = 0 •
^ ^ a / 3 » + ^ ^ V + PSX* = 0 .
(7.37)
Equations (7.37) are objective and this suggests that /* and L* are physical quantities for which we must give constitutive equations. On the other hand, (7.35) looks non objective owing to the presence of r x / and rx F ; however, it is easy to recognize that (7.30) implies the objectivity of (7.35). In fact, under a change of frame of reference I -* V given by
i ' = x o (0 + x ,
(xn = x'0 + Q)x^,
Q'jhiQ1 )] = &))
due to the objectivity of / , / * , . . . we have /' = / ,
r ' = r , . . . , r ' = x0 + :
177
CHAPTER 7. CRYSTAL GROWTH
and thus from (7.30) and (7.35) it follows that
p(l*' + r'x / ' ) dv + [ ps(L*' + r' x F " ) d
p(l* + rxf)
dv+
ps(L* + r x F) da +
C
ee
psF da + £
L' dl r
s£
+ x0x(\pfdv x M pf dv + +
L' dl
# d/) = 0 . i
Moreover, when applied to the balance of energy, eqs. (7.24)-(7.25) yield
pe-tr(t®Vv)
+ V-(h + n1q) = p(e-f-v)
,
in C - E ,
(7.38)
^ ' - ^ ■ T - V ^ +V - ^ + ^ g J
-[/»(£(*- - ^ ) 2 + e _ ^ ) C f
+ n.(f.(v_v)_ft_/iig)] =
^ _ F . y ) , onE-T,
r ■[nx [(C- -V)p,( I ( T ^ - C ) 2 + £ ) + T- (V--Oon
= - ( § - # C), pe dv+
psE da + E
H~vM'
(7.39)
g dl = 0 . r
we get
w
'j = v[},t}
-VulsW
r,
On the other hand, if we introduce the notation
d
K
= ^jhw
>
178
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
tr(t ® Vt>) = t>vjt i = t^d4j
+ < M W y = t^di}
+ t^%hwh
.
(7.40)
Similarly, on the surface we have
VSV = a"®V,a
= (yp.
a
- Vnbaf})aa ® a? + (V B| „ + 6 a / 9 v V ® n
= D+W , where D = aapaa ® aP + ffa{aa ®n + n®aa) , a
a
W = Wafia ®aP
a
+ (Ta{a ®n-n®a )
(7.41)
,
and a
ocH -
V
(fl; a) - Vnbal3
"a^hfY^ W
afi
'
(7.42)
+ b^Vfi), =
V
[0,a]-
We introduce the surface curl of V by
Q = V,xV = ««xK, a = ( ^ ; « - V n ) « " ^ + ( f n , a + *a/^ / 3 ) a" x n
=
> / 3
whence it follows that
«„ = f"%; « = ^ ^ - «a = 2f *% . or
^
= K/3 f i "' ' - = 5 ^
(7-43)
CHAPTER 7. CRYSTAL GROWTH
179
It is then straightforward to verify that the difference of velocities of two points on E which are apart by a vector dr is given by dV = dr-D+[ dr-D+[^Q±n + nQaaaaaa]xdr ]xdr nn nn
(7.44) (7.44)
.
Hence the angular velocity on the surface is the vector n = nQaaaaaa + Q Q nnnnn. .
(7.45) (7.45;
It is now easy to verify that (7.41)-(7.43) imply {al3) aa ■ • TT ■■ V, V,aa = =TT{a0) aaa/3a0 ++ T^W^ T^a^Wap + + 2T 2TaaD Daa
=
Tt«n*afi+ica$T"enn+tafF°aP. (a
= T ^a0
+ i ca0r*iin
+ tal?*&
.
(7.46) (7.46)
By recalling (7.37), (7.40), (7.46) we get By recalling (7.37), (7.40), (7.46) we get
tr(t V«) = fi^dy -pl*-u>,
(7.47)
{ij)
(7.47)
tr(t $ V«) = t di;j -pl*-u>, aa
■T.V, aa .T.V, a a a
a .T.V,a
t
= = =
Jl(l
T^a -PsL*.n. ■ n. a0-Ps a0L* a/3T^a 1
T^aa0-PsL*.n.
Introducing (7.47) into (7.38) we find the final form of the local energy balance equations pe-t-{i3) dtJtJ pe t^d
) + V-(h V • (A++fiHiqq) +
= i nC C - E ,, = p(e-f-v-r-uj) p(e - / » - ? * ■ ») = = pr pe* , in
(7.48)
180
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
P,E'-
j(a
B )
° a ^ s ■(*» + P i . O +ee-E)U n-{t-{v-V)-h-n -E)U ++n •(*•(* -F)-ft -Plfl)] + iq)}
-lp{\{v-Vf -[/»( \{v-Vf = PpSs(E-F (E-F-V-L*-Sl) psE* , oonnSS- -•r, r, ■V - L* ■ Q) = =/>,£*, T.lnx[(C-V)Ps(±(V-C)2
+ E) + T-(V-C)-hs-p.uqs}}
= -(e-#-c)= -s*,
onr
.
Finally, (7.30) becomes
p(e* + f-v + l* ■ u) dv + ps(E* + F-V +
L*-n)da
(&* + <J-C)dl = Q.
(7.49)
The quantities e*, E* and 8* are objective in view of the condition (7.49). In fact, under a change of reference we have
L(e'* + f'-v' + r*-L>') dv = \p(e* + f-v + l*-u) dv + x0- \fdv + t
e
c
{rxf
+ l*)dl,
c
since t/ = t> + i + u > x r , LJ' — u+w . By applying the same considerations to the other terms of (7.49) and adding the results, we derive the invariance of (7.49) as a consequence of (7.30), (7.35).
CHAPTER 7. CRYSTAL GROWTH
181
7.5 Nonlocal reduced dissipation inequalities.
We postulate that the following well-established inequality is valid for all admissible processes which take place in the whole system d_( psdv+ ps dv + ■4j( C
da)> psSS d(A>
±h-Nda\h-N da-
aC
1£
\h \h,.*zdl s-Vlidl s dl
(7.50)
as as
(see (3.37)). If we localize (7.50) we obtain
(see (3.37)). If we localize (7.50) we obtain
A» + V - gjf) = ps+V-(£j
/» ,
in C - E inC-E
= ps ,
(7.51) (7.51)
«v
-lp(s-S)U-ih-nj = p,S psS ,, on E - T , - M « - S ) t f - 1 A.n] = PSS' + VS( *) & PsS' + Vs-fy)-lp(s-S)U-ih-n} = psS , o n E - T ,
r-[»x[(C-K)p,5-iy =
-» , r.[nx[(C-V)pa5-lA,]]= -5 ,
on T , on I \
where ps dv \p$ dv++
cC
S£z
psSS da da++ J* dd// >> 00. .
(7.52)
ir
However we will now show that (7.51), (7.52) can be transformed to a form which is more suitable for applications. Let us partition localization residuals arbitrarily in such a way that
= s S+ + Ss* * ,, sS =
S5 =§ = 5 ++ §* 5* ,,
f£ ==?? ++?* ?*, ,
where S s > 00 ,
in C - E ,,
55>>0 0 , , on o nEE-- T ,
t? >0 > 0 ,,
on on TT ..
182
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let us now denote the left-hand side of (7.52) by R(t). Of course depends on the thermodynamically
process and the only
R(t)
information
available about its properties is that it is non negative for all times. Since s, S, 1 are chosen to be positive, we can suppose that
psS da+ \$ dl = R(t)>0
ps dv + C
S
.
r
For instance, a possible choice is provided by s , S , !f subject only to the constraint sMc + SM^ + f/ = R(t) > 0 , where M*> and M s are the masses of C and / is the length of T. Hence relations (7.51) can be expressed in the form
ps+V-(£\-ps*
PSS' + Vs-fy)-lp(s
r-lnx[(C
= ps > 0 ,
-S)U -± h-nj-
-V)PsS
inC-E
(7.53)
psS* = psS >0 ,
-j: hs]} + r = - ? < 0 ,
onE-T,
onT
We call the fields s*, S*, !f* as localization residuals. They satisfy the global condition ps* dv +
psS* da +
r di = o.
(7.54)
If we use (7.38) in (7.53) to eliminate V • h and V s • h , we arrive at the
183
CHAPTER 7. CRYSTAL GROWTH
nonlocal form of the reduced dissipation inequalities
(i])
-p(^ -p(i>++s0)80)+ + t(ij)tdij-'V-(ii dtj 1q)-^h-V9 - V • Oilff)
- ps(V
+ S6> -
- 1 A • +V0 p^*>O, + p#* > 0in i ,n C-E C - E ,,
HgV's)
+ T^Da(i
+ [p[ \ (« - V? + g - * - nls(u
-
9s
(7.55)
■ VsMls
- vs))V + (Pls -fijq-
+ [n-t-(»-V)]+/),**-ift,.V,»>0 ,
n]
on E - T
T.l S-±h 9*== - ? < 0 ,, s}]) ^ - U j ] + J* r . n[x[(C-V)p n x [ ( aC -V
o nT T, on
where g = ip + p/p is the specific Gibbs potential and we have put V>* = e * - 6 > r ,
$* =
E*-6S*.
W e could here choose a class of constitutive equations for all the
fields
which appear in our balance equations and try to derive thermodynamically restrictions
for
these
constitutive
equations
including
the
localization
residuals. W e are not interested in developing so wide a nonlocal theory of phase transitions; on the contrary, we prefer to treat a restricted class of materials in order to test the effectiveness of this approach.
7.6 Preliminary considerations on crystal growth.
In this section we want to recall some fundamental aspects of crystal growth to have in our hands elements to control the results of our model. It
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
184
is well known t h a t a crystal of a substance C can be produced i) in the melt or vapor of C; ii) in a solution containing C as a constituent; Hi) by the presence of impurities in the melt, vapor or solution. It is easy to understand t h a t the first case is very difficult to realize experimentally. On the contrary, the
second
one
is
much
more
practicable
and
allows
a
control
of
crystallization process. The third is almost spontaneous and consequently it is very difficult to describe and govern. Whatever is the environment in which the crystal grows, the process always starts from a germ.
This is an agglomerate of few hundred
or
thousand molecules which are temporarily linked to each other as a result of statistical fluctuations. When the temperature and pressure are given, the destiny of the germ depends on the ratio between its surface and volume. If this ratio is below a critical value, the agglomerate dissolves; in the opposite case, it grows up and attains a configuration which is governed by Wulff rule (see sections 7.1, 7.2). Starting from a germ, the crystal increases under suitable temperature and pressure conditions. From now on we suppose that these external conditions guarantee the regularity of the growth process so that
the
crystal
faces
are
planar
and
advance
remaining
parallel
to
themselves. Two fundamental theories have been proposed to interpret the crystal growth. Classical
theory.
According to this theory the growth is due to three
main factors: i) Adsorption:
The atoms and molecules on the crystal surface
are
captured, from the surrounding phase, by the molecules already stored in the
CHAPTER 7. CRYSTAL GROWTH
185
crystal. Some of these new collected molecules, owing to thermal oscillations,
Fig.24
go back to the surrounding phase after a mean time which depend on the substance and varies from 10
sec. to 10
centuries. Other molecules, due
to the greater number of interactions they are subjected to, definitively fit into the steps which are present on the crystal face (fig. 24). ii)
Surface
diffusion
: the captured molecules migrate toward the steps
(fig. 24). Hi)
Surface
nucleaiion
: to allow the face growth after a molecular layer
has filled the face, it is necessary the creation of a two-dimensional nucleus which plays the part of a crystal germ for the new layer.
186-
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Fig.25
Fig.26
187
CHAPTER 7. CRYSTAL GROWTH
Burton, Cabrera and Frank [58] showed that the probability that a new two-dimensional germ is created on a face has appreciable values only if the vapor or liquid supersaturation rj = (p - p0)/p0,
where p is the pressure in
the vapour or fluid and p 0 the equilibrium pressure, is greater than 25%. On the other hand, it is possible to obtain significant growth velocities for values of 77 ~ 1%. This gap between experimental and theoretical d a t a requires a new model to describe crystal growth. Dislocation
growth. This theory foresees the presence of dislocations (figs.
25, 26) which generate one or more steps on the crystal faces whose height is a few molecular layers. These steps are stable since the molecules are stored on the face without modifying the distribution of the steps. During the growth, which is a consequence of the installation of new molecules along one step, the latter makes a spiral around the dislocation. The
mentioned
Authors evaluated the number of dislocations, which could be present in a crystal, and very high values were found. For instance, in 1 cm
there could
be dislocations for a total length of about 10 cm. These growth spirals can be observed by interferencial and phase contrast microscopy.
7.7
A
mathematical
model of crystal growth in a
binary
nonreacting mixture.
In this section we apply the general nonlocal balance laws to a particular
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
188
continuous system with an interface (CC,C, E) which is able to describe the main characteristic of crystal growth. More precisely, from now on we suppose that: i) the crystal phase Cc is rigid and at rest whereas the phase C is filled with a perfect classical binary mixture; consequently, the unknown stress tensor tc in the crystal has to be determined by the evolution equations; it) the phase Cc is a convex polyhedron so that the interface
E = 0ec=U',-.
J = l,2,...,n,
(7.56)
t' = 1
is the union of n planar faces ai of the crystal; the broken line
is the union of n planar faces a^ of the crystal; the broken line m
r =PUm= i/„, T = U L,
p == 1,2.....,m ,
(7.57)
p = l,2,...,m,
(7.57)
which represents the union of the edges / of Cc , is a discontinuity line for the fields defined on E. The only physical attributes of each ai are represented by a surface stress tensor Ti and a surface energy density E^ Moreover, on each cr^ , the surface stress tensor Ti is a function of the values on
189
CHAPTER 7. CRYSTAL GROWTH
iv) the localization residuals are absent in the bulk phases but are present on £ and T ; in other words, the nonlocal effects are related only to surface phenomena (see section 7.3). Finally, we neglect the external body and surface forces. When we write all the equations (7.26)-(7.55) in these hypotheses we get
p +pV-v=0,
in C ,
(7.58)
pv + V • q = 0 , pi, = - V p , pe = -V-(h
+ niq) ,
pc = const. ,
in Cc ,
(7.59)
V-< r = 0 , t= Pckc= -
[[pU] 0, PU] = Q,
tT, V
9c .
oonnEE--TT,,
(7.60; (7.60)
lpUv]-q-n IpUvJ-q-n^O, = 0, -Vs-T-lpvU rxF=L 2
^-lp(±v
+ n-q = F , ,
+ e)U + n-t-v-n-(h
+ ti1q)} = E ,
where in C , e is the partial time derivative since the solid phase is at rest. Finally, we have (see (1.64)'):
190
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
r - [ r » x T ] = -{v-T}=
-9,
on T ,
(7.61)
rx§ = £ , f [ n x £ ; C ] = -{EC-v}=
-I,
r - [ n x 5 C ] = -{£C-i/} =
-If.
Here C is the velocity of T, v = rxn
and the localization residuals satisfy the
integral conditions ,Fdcr + f 5 ( f l = 0 , E f rxF
(7.62)
r x 5 d / = fl,
da + r
E
LB
J E
r S da +
f d/ = 0 .
We remark that we have neglected the dissipation on the edges. We are interested in those crystal growth processes in which each face moves remaining planar and parallel to itself so that the final result is a regular crystal. We assume, to guarantee this kind of evolution, in agreement with experimental procedures, that: i) all the fields are uniform on each crystal face; ii) the stress in the crystal reduces to a pressure; Hi) the growth velocities are very low ( ~ 1 0 - 6 m/sec.) so that the
191
CHAPTER 7. CRYSTAL GROWTH
process lasts many minutes or some hours. This last circumstance permits to neglect some terms in the equations (7.58)-(7.62). In fact, in ordinary conditions of pressure ( ~ 10 5 N / m ) and crystal dimensions (10 ~~ -r 10
m), by using a nondimensional analysis
(see sections 5.2, 5.3), it is simple to derive that the pressure is uniform in the mixture and the terms pvll,
pvU/2
are negligible with respect to the
others. Moreover, if O is an internal point to the convex crystal and hi the distance between O and the face ai we have
C
ij{p)H(p)-nnii=Cni = Cn. = ilhii> '
(7.63) (7-63)
where C- ■, •> is the velocity of the p-th edge /,-•»_% between the faces a., cr ■; finally, the T h o m a s derivative vanishes in (7.60) 5 since E depends only on n and n remains parallel to itself during the whole process. When we take into account (7.8) as well as all the previous remarks, we can write equations (7.58)-(7.61) as follows
p + pV-v
= 0,
in C ,
pi/ + V ■ q = 0 , P = Pe>
pe = - V - ( A + / * , « ) ,
(7.64)
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
192
pc = const. , tc=
(7.65)
in Cc ,
-Pe{t)I,
pch= - V A > vn.=—p-^h^phi
,
(7.66)
on or,. , i = 1,2,....n ,
9"wi = / ' c ( y - 1 ) ^ . T, = const. ,
F, = Ftn{ ,
[p] = F, ,
rxF,.=2,., P c ^ +^ l + ^ C i - ^ ^ - W - " , - - ^ - , (7.67)
-(Ty»i + ^ » i + ^ o - ) = - ( % " ; + V » i + % * - ) , on / t j ( p ) ,
p = 1,2,..., m ,
-(V^A'}=-%-' r
here where • T > • ni - va ■ Tj ■ nj
*u = ^"iji^ji «?y = CSC
^
2
a i J [# t J
«i =
E
j
ij]
.
(7.68)
• n,- - # ( i • Ti^ COBOy] ,
= («/ IJ -T 1 + „ J 1 - T J ) - r l J , A
cosa
CSCQ
ij
- Ei
^■= # , r r . i l coia
i]
■
To derive (7.68) 1 . 4 it is sufficient to represent both sides of (7.61) 1 in terms of the basis (n,-, n-, T-) which we fix along the edge / • ■. Equation (7.68) 5 is obtained by (7.8), which allows to express vi ■ and u •■ in (7.61) 3 in terms of the aforesaid basis, and taking advantage of (7.63). For the sake of simplicity, in the sequel we suppose that
193
CHAPTER 7. CRYSTAL GROWTH «P. • = <J'- • = 0 .
To obtain the field equations describing the crystal evolution it remains to introduce constitutive equations, initial and boundary conditions. First of all, (7.55) lead to the following reduced dissipation inequalities
-p(V> + s f l ) - p V - t > - V " ( / i 1 f ) - i f t - V 0 > O ,
-p(^c
+
sJ)-lhc.V9>0,
PclV' + § + ^ 1 ( l - ^ ) ] A l
+
in C ,
inCc,
*I + F t / l l > 0 ,
(7.69) (7.70)
on<7 t ,
(7.71)
where if) = e-6s is the specific free energy and $,- = Ei - 9iSi . In the bulk phases, according to the classical mixture theory and the reduced dissipation inequalities (7.69)-(7.70), we adopt constitutive equations of the form (see section 6.2):
iPc = iPc(6), , ,/ m lP = 1P(P,1S,0) ,
s = - ^ > dtp S=—QQ,
(7.72)
hc=-k-Ve, 2H P = P-T^,
dip V1=-fo,
h= -kve, q = q(p,v,e,vp,vv,ve), ^ - v ^ - ^ w f ^ o , (7.73) where k is the thermal conductivity tensor in the crystal and k(p,v,9)
is the
thermal conductivity of mixture. It is evident that we must also consider
194
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
constitutive equations for F t , Ei , ?,- ■ , 84-j but we discuss about t h e m in the next sections.
7.8 On the crystal equilibrium.
At equilibrium the system (7.64)-(7.67) reduces to the following set of equations p(p,0,v)
=
inC,
Pe,
pc = const. ,
(7.74)
in C r ,
[ p ] = Fi > on o-,- , ?' = l , . . . , n ,
- (**««,• + V » i ) = - @Hni + V 1 * ) '
on
W ) 'P = 1 - " m '
to which we have to add the integral conditions (7.62) 1 Now an equilibrium configuration
2
(7J5)
.
of the crystal in the mixture is
characterized by the uniform values of 8, pe, p, v and pc as well as by the orientation n, of the normal to the faces ai and their distances hi from a point O internal to the crystal. When we assign constitutive equations for F-, 9- ■ as functions of the aforesaid variables, (7.74), (7.75) and (7.62) x
2
represent a system o f 2 + n + m + 2 = 4 + n + m equations with 5 + 3n + n + 2n = 5 + 6n unknowns pe ,pc , p ,0 ,v , Ti ,h , n , , where h = (hv...,/in).
In
the next section we show that the reduced dissipation inequality on E allows one to derive another condition on the faces in the same unknowns. This
195
CHAPTER 7. CRYSTAL GROWTH
means
that,
in general,
t h e equilibrium
configuration
is n o t uniquely
determined. However, b y taking into account the crystal symmetries, we could determine some possible configurations which are compatible with these symmetries and the equilibrium conditions. For instance, for a cubic crystal, we have T , = yj where 7 is a surface tension which does not depend on the face. T h e n we can ask whether a cube represents a possible equilibrium configuration for the crystal. In these hypotheses we have
p{p,vt8)
=
Pe,
pc = const. , [ P ] =Fi 1
o na
i .
-7(n,. + n i ) = - 5 ( « , . + n i ) ,
inC, in C r , i = 1,...,6 ,
on l.j{p) ,
„A
p = l,...,12,
), - * * 21 ("* ++n»;),
PFEi"i=-I*%.P(n P i 1
(7.76)
~ 12 12 ,. .. rr s,_ d
(7.77)
(7 (7.78)
i
r .11
x
1
'« where / is the length of the cube edge, r is the position vector with respect to the center O of the cube and rs is its projection on the crystal face. It is clear that both sides of (7.78) 1 vanish separately for any choice of /, F, 7 . Moreover, the integral on the left hand side of (7.78) 2 represents the center of mass of cri so t h a t rs = 0 and each term of the summation is zero. Similarly, the integral on the right hand side of (7.78) 2 denotes the center of mass of the edge /
and consequently the right hand side also vanishes. We can
196
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
conclude that system (7.76)-(7.77) reduces to four equations with unknowns pe,pc,p,v,0,y,
h = 1/2. If we add to those the other conditions on the crystal
boundary, which we derive from thermodynamics, we attain the conclusion that the variance of these cubic equilibrium configurations is two. Similar considerations can be repeated for the symmetric polyhedrons. We note that the only a priori information about the form of the constitutive equations of Fi and ^i ■ are represented by the restrictions which are imposed by the crystal symmetries. However, by experiment we have that [p] = 0 for large crystals so that lam F- - 0 .
(7.79)
This, in turn, implies that the equilibrium configurations of large crystals have to satisfy the equations p{P,6,v) = pe, Pc = Pe>
in
inC, c
(7.80)
c
so that these configurations are arbitrary in form and have again variance two. In the case of small crystals, the integral condition (7.62)j can be written as
t ,•*>.•*,• = - E P@iilp)*i +
*ji(P)ni)lij{p)
p
= - s Pc*ij(p)ni+*
jnp)njhj(Py
The last two terms of this equation contain the same number of terms as pairs of adjacent faces. Therefore these terms can be combined as it is shown
CHAPTER 7. CRYSTAL GROWTH
197
in the relation
t i?Wi = ~ E i E ■ IH*^ = - t i E 'i hFi,**
(7-81)
where the notation E i means that the summation is extended only to the edges which form the boundary of or.. When we suppose that i) the condition (7.81) is satisfied as a consequence of the following hypothesis $Wi
= ~ E ■ hpiFi = - E • ijij^i,
(7.82)
H) Tt = 7,-J». it is easy to verify that (7.74) 3 and (7.81) lead to the relation
VC~V = \ E j V . j ' where ij = 7y csc Q j i - 7,- cota t J .
This relation coincides with the Gibbs rule (7.16)3 when j i = Ei (see remark it) in section 7.1). The following considerations supply a justification of hypothesis (7.82). Let us regard the crystal surface as a set of planes which are constrained to remain parallel to themselves during the phase transitions and in contact along the crystal edges. The crystal faces and the edges interact with the
198
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
volume particles which are contiguous to them and these interactions are represented by the localization residuals of surface and edge, respectively. Consequently, the edges are subjected to the forces 5 , and the crystal faces are subjected to the parallel and uniform forces F,n t , which are independent of the face. We have no phase transition if and only if each face is at equilibrium. Now, it is easy to recognize that the face ai is at equilibrium if and only if (7.82) is satisfied. In fact, let us consider a virtual displacement owing to which the crystal reaches a new configuration in which only the face ai is translated by the amount 6h^ along the normal n,-; this implies that Sh ■ = 0, j ^ i, whereas the edges of the contiguous faces move to remain in contact with
iW\=-£;',A-•*•.;>
( 7 - 83 )
where <5r-■ denotes the aforesaid displacement of the edge /•• and therefore verifies the condition Sr-■ ■ n = 0. Since it is straightforward to verify that
Sr
ij =
csc2
% K -
cosa
«jnj).
equation (7.83) reduces to (7.82) when we take into account (7.68) 2 . Hence, (7.82) could represent a restriction to be satisfied by the constitutive equations of F-•, 'i- ■.
199
CHAPTER 7. CRYSTAL GROWTH
7.9
The
energy equation a n d t h e reduced dissipation inequality.
If we introduce the latent heat
A= [ 6 + ^ 1 + ^(1-1/),
(7.84)
the energy equation (7.66) 5 becomes
AA, + £ i
=
[fcJ.n,.
(7.85)
As is well known, the latent heat is the amount of heat which the unit volume of the substance must adsorb to pass from the solid state to fluid or vapor state. This adsorbed heat generates a j u m p of heat flux. However, (7.85) shows t h a t only a part of the heat flux j u m p is used as adsorbed heat because of the presence of the energy localization residual E'-. We have to justify the presence of this additional term in (7.85). First of all, we remark that in the absence of E^ the growth velocity of the crystal faces would depend on temperature 9, concentration v, external pressure pe as well as on the orientation of the face owing to the heat conductivity tensor, which in the crystal is not isotropic. This dependence generates differences among the velocities
of
crystal
faces
which
are
too
small
compared
with
the
experimental velocities. Therefore, we could think of assigning to Ei the task to describe this strong dependence of the velocity on the crystal face. In order to understand the meaning of E^
we take into account (7.61) 3 to write
global condition (7.62) 3 in the following way [59, 60, 61]:
200
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
n
m
(7-86)
E i E, *t = E tj(p) (E,Vij + V / . 0 •
The left hand side is the total excess heat adsorbed by the whole crystal surface during the growth and it is equal to the total energy associated to the creation of new surface parts per unit time. This consideration supplies a remarkable meaning of Ei but, at the same time, would lead to the wrong conclusion that the presence of Ei is meaningful only for small crystals for which the energy per unit area necessary to increase the surface of the faces is comparable with the energy adsorbed by the unit volume in a phase transition. In fact, (7.86) is a global not local assertion about E±. What follows is to make this point clear. The integral condition (7.62)3, when we take into account (7.67) 3 and repeat the considerations relative to (7.81), becomes
£< V.-=-E,-E'V.-i A .-« 1
1
(7-87)
3
so that we derive f
.-=4E'Wi + ^
(7-88)
3
E,-^i«r,- = 0.
(7.89)
With these notations the energy balance equation (7.85) 5 can be written
( 4 E '
'AijLij), ^ = [A] • n, - E\. 3
(7.90)
201
CHAPTER 7. CRYSTAL GROWTH
It is easy to verify that (7.86) reduces to (7.88), in which we put E\ — 0, for a growth process corresponding to hi ^ 0, h = 0, j / i. In other words, to suppose
E'; = 0 is equivalent to say that the localization residuals Ei of the
face <Xj depends only on the growth
velocity of
this
assumption only for small crystals, the quantities E\ have to satisfy the property Urn E'- = 0
(7.91)
and therefore it describes the growth process for large crystals since it is obvious t h a t when E\ = 0 .Urn
^ - ^ E ^ / , / , - 0 .
| h | - » oo
«
(7.92)
j
It is possible to give a physical uinterpretation for E^. In section 7.6 we have already noted t h a t the theory which better fits the experiments on crystal growth is t h a t of spiral growth. According to this theory, a condensation nucleus of suitable radius is formed at a point O of the face and it gives rise, as a stable dislocation, to the expanding spiral on the face. T h a t suggests the introduction of a surface energy flux j associated with this expansion and depending on the face. By applying the Gauss theorem to the whole crystal with planar faces we obtain n
E, i
m * V s -j da-- ? P
■
{ii'Vij +
i / J *j (p)
T h e previous expression can be put into the form
iyVjddl-^.
202
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY n
=o i dcr E * ) * < ) = o, E,(^(jv ' s -i^-E'ji i -?'/'' *r*^0^ 3
'y '.«•
where the notations were previously introduced. Consequently, the quantities
^ ^ ( [ V ivA r - E ' U - hi ^■tvaf )dl,) s -jrf
(7.93)
hi
associated with the face
e,h,k).
It remains now to consider the reduced dissipation inequality (7.71). If we extend the decomposition- (7.88) to the entropy localizations V •""y
s,= -ilZ'Bijtijhi + Fi, SJ. £'s;-^=o, : ' % '« A,;+ '
3
1
and recall that $ ; =
E--6SJ
3 3
, we can write (7.71) as follows
p E' % PcM+Tl - i ^- ^-) -h^ S - * 'Vt Jy)) ^* + i +**; }>>o0,. c ( ^ + ^ ]++^ M
(7.94) (7-94)
where DiS =
, VAij-*ii-0B -V ij
£),■*; *,- = o.
The left hand side of (7.94) can be regarded as a function / • of h- (i = 1,..., n) which has a minimum when h; = 0 (i = l,.,.,n) so that at equilibrium we have
r«£n oh^ y
h
==0 0
= /»c(W-►?]- f / i j ( l - - " ) ) - -E f j
£»■■
J
,a*j J
^ dhi \
L )
= Q=
0
i
CHAPTER 7. CRYSTAL GROWTH
203
± i, (wl \dh-h.n =(#)■ V a / , , 4 - n =•■ i'*'• v
dhi
J
h
= Q
v a/ij ih _ o
$i Ai0»,v ,.,n), $J = ■**(/ >, M1fl)h. ) \ - ++*J' *J' (,-(i == !,.. l,.-.,n),
(7.95) (7.95)
where the functions ty" are at least of the second order with respect to h^ Consequently, when we insert (7.95) into (7.94) we obtain the following j u m p conditions on the crystal faces
PC(N> + ^ 1 + ^ i U - ^ ) ) - £ % ' , , +A, = 0 ,
(7.96)
3
as well as the thermodynamical restrictions on the functions
$'.' %'{ >0 > 0 .. In the sequel we also accept (7.96) in dynamical conditions since we consider here only very slow processes. In order to derive an interesting consequence of (7.96), we introduce the notation a M to denote the value of the quantity a for a large crystal. In these conditions (7.96) leads to the relation
P ^] + M / i i*( l -~- ""))oo PcW ) ) « , + ( A oo), = ^oc(Pe> o c ) , = 0° ,. ^oo(Pe> 0oo^oo- O "oo) + ( A oo)t c(lV> + £
(7.97) (7.97)
which allows to evaluate the equilibrium temperature 6^ as a function of pe and v. But at equilibrium the temperature is uniform and this imposes that the quantities (A )■ do not depend on the crystal face. Moreover, we can put
204
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
(^oo)i = °
since
in
tne
absence of this term in (7.97) we obtain values of
equilibrium temperature which fit the experimental results. On the other hand, to within terms of the 2nd order, we have A:
=mj°- •o=
-•Boo =
(e- - ' « , ) .
and thus (7.96) can be written as 6- -*oo =
*oo " S oOO o
[
1 *
E 'Di3 hi-- A J . J
j
This relation shows that the quantities within the brackets must not depend on the face and that, for a small crystal, the equilibrium temperature 8 could be smaller than 6^ (supercooling).
7.10
The free boundary value problem describing the crystal
growth.
The evolution of a crystal is described by the system (7.64)-(7.67) which is a system of partial differential equations in the unknowns p, v, 6, pc, v. A set of reasonable boundary conditions associated with this system can be derived by supposing that the mixture is contained in a vessel. The walls which are in contact with the fluid phase will be denoted by S. If the free
205
CHAPTER 7. CRYSTAL GROWTH
surface of the mixture is 5 ' and the crystal is far apart from the walls, we can give the following boundary conditions
onS ,
6, vn, q-n = 0, Q, P = Pe> q-n-0,
(7.98)
on S ' ,
0, vn, q • n, Pc,
on E.
We remark t h a t the nonmaterial character of E allows us to assign on it a velocity vn and a pressure. Now, the d a t a (7.85) 3 are given in terms of j u m p conditions (7.66)j
2
3
on
the unknown surface Sigma whose evolution is in
turn described by (7.66) 5 (generalized Stefan's problem). Finally, another condition on E is represented by the inequality (7.71) which we assume to be satisfied in the form (7.96) (slow processes), in what follows. W e are now able to give the precise mathematical formulation of the boundary problem which describes the crystal growth. We have to find the fields p, v, v, 0 in C and the field 9 and the quantities hi in C c such that
p + pV-v
= 0,
in C ,
(7.99)
pv + V • q = 0 , pe = -V-ih
PcK=-V-qc,
jB. = ^ i
f
H j 5 i j ,
+ fitf) ,
inCc,
on a,-, i = l , 2 , . . M n ,
(7.100)
(7.101)
206
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
qn (A- 1
= p c ( j / - l ) h{ ,
£''4i ' « K = [*]• n, - ^ . ) , ■
'
3
pc(W+^] + M i - ^ ) ) + E * ^ ^ + ^ = oi
(7.102)
In the previous equations the functions A-;' s are defined by (7.68) 4 ,
L r csca csca LL cota
D D
i.,
(7.103) (7.103)
where the Li are, together with the quantities A-, given constitutive equations of the variables p,v,9,E
■,hi and the E\ 's are functions of p,i/,6,E
-,/i ,A .
Moreover, E\ and A ■ satisfy the global conditions
E i ^ " i = 0, l
E , \ ^ T = o,
E . A . ^ ^t ^ O . il
(7.104)
W e have also defined (see [44])
« = 5 £ '(ftj c s c a ^ - &,. c o t a ^ ) li3 , i
= *«i = hi
V
V
- 11
ccsca. sca
o t ot * , \3-i --11 ccoifi ^ i -_- 1l,. J' j - -hftii ccsca M O f*^( «c *o /t ?^, -_. -l1f . ii J++ ccot/?^ ^ - . , ^. +j +x )x)
csca csc csc/^+ h i[coto,-, ( c oi_thj^ . . , , -+r C + ** i; + O t ^h}. i ++ 1i)) cotf3 lii{cotp + 1i csca-i ,,ii + + 1l Pj,j■ j ++ 1l + hi[eoto -li
csc -- cota cotaj csc/? i : ,.i ++ iJ' i, , jj --1 ^ i _- -ii,-i- -j "c-o ct o t,a^; ,,-y++ ji csc/?^ J, - l cscf3j
CHAPTER 7. CRYSTAL GROWTH
T T w h e r e cos/? cos/?^ j■++ i~ where 1 = r ij' . - i ' rii , j+1 i +l
207
an( a n^ d Trij
denotes the the unit unit vector vector along along the the »i denotes
edge formed by the faces ai and a •. This problem is very difficult but it is possible to simplify it substantially if we are dealing with large crystals (i.e. having linear dimensions (i.e . crystals having greater than 10 _ "' m). In this case we have
p+pVv=0,
in C ,
(7.105)
pi/ + V • q = 0 , pe = -V*(h
+ nxq) ,
PA=-V-AC,
vn,=^frhl
= l3hl,
inCc,
on*,., t = l , 2 , . . . , n ,
V l n q-n{ = {v-l)h = pcPcV - U i i, « • « , ■
Xhtz
= [*]•"*- ^ > [^+7?I+Pi(i-»') = 0.
(7.106)
(7.107)
Chapter 8
SYSTEMS WITH INTERFACES AND FERROELECTRICITY 8.1 A b r i e f s u r v e y of f e r r o e l e c t r i c i t y .
In a dielectric crystal C the polarization per unit volume P is a function of the external electric field E
which depends on the nature of the substance.
When the electric field is very strong, P tends to a constant value and we say t h a t the dielectric is close to the saturation because almost all the molecular dipoles have been oriented by E. In particular, P is parallel to E if C is isotropic. A dielectric crystal C is said to be ferroelectric if very high values of polarization, in particular the value which corresponds to the saturation, are obtained by applying very low electric fields (curve 1 in fig. 28). Moreover, if we reduce E after the saturation has been reached, the polarization
assumes
the values corresponding to the curve 2 and reduces to zero when we apply an opposite field -Ec
which is said to be a coercive field. When we again
increase the values of E, we obtain the values corresponding to curve 3. The set of curves 1, 2, 3 represents an hysteresis loop.
208
209
CHAPTER 8. FERROELECTRICITY
fig.27
This anomalous behaviour of ferroelectric crystals has to be found in the domain structure of polarization. More precisely, even in the absence of external electric fields, the polarization in a ferroelectric crystal is represented by a vectorial field P = P0p,
where P0 is a characteristic scalar of the crystal
and p is a unit vector field. This latter is piecewise constant in almost the whole crystal C except in thin layers across which it varies from one constant direction to another. In general, the regions Da, a = 1,2,...,n, in which p is constant (Weiss domains)
have at least one microscopic dimension (a few
hundred microns) whereas the others are comparable to those of the crystal. Moreover, the aforesaid transition layers (domain
walls) have a thickness of a
few microns. T h e theoretical justification of this complex distribution is to be found in the combined effects of three factors: the electric energy due to
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
210
the anisotropy of the crystal, the dipole-dipole exchange q u a n t u m forces and the form of the crystal. The presence of an external field can only modify the weight of the effect of these three factors on the distribution of the domains. F r o m a classical viewpoint one takes the first two factors into account by associating with the crystal an energy e(p,Vp)
per unit volume in which the
contribution due to the presence of V p is significant only for very large values of || V p || . Then the evolution equations of the crystal are so complex t h a t it is very difficult to deduce some general results from them and very few information on the distributions of domains, on their evolution or on the energy contained in a transition layer was obtained only by resorting to rather drastic approximations [62-73]. In order to obtain a description sufficiently accurate and also to simplify the situation illustrated above, in [74, 75] it was suggested t h a t the domain walls are to be substituted by surfaces of discontinuity Sb, 6 = 1,2,...,3 of p and to take into account the energy contained in a transition layer with a surface energy E(n,p)
depending a priori on the orientation of the normal
n with respect to crystallographic axes and on the constant vectors p
in
two domains adjacent to the domain wall under consideration. However, in addition to this surface energy we still go on to consider an energy per unit volume of the crystal depending on V p besides p. In fact, except for very particular forms of C related to the symmetry class of C, the satisfaction of the boundary conditions on <9C may require the formation of small regions adherent to the external walls where the polarization field is not uniform. The scheme we are adopting suggests to use again the model of continua with interfaces to describe the evolution of a ferroelectric crystal. In next sections
CHAPTER 8. FERROELECTRICITY
211
we show t h a t this approach is fruitful.
8.2 Maxwell's equations for a moving continuous system.
Let C be a charged and electrically conducting continuous system which moves in the inertial frame I. We denote by V C C any material volume and by I f c C any material surface; moreover, we suppose that there exists a moving (not necessarily material) surface E C C across which the volume fields undergo discontinuities. Then the Maxwell equations for C become (see for instance [76] and the general balance equations (3.1), (3.5))
d dt
B-Ndcr=BN da=
jj
j
j-\D-Nd(j= D Nda=
Ttl ' f
-
1
(E+VXB)-T (E+VXB)-T
(8.1) (8.1)
dl ,
af
\df
(H-vxD)-Tdl(B-VXD)-T
JN daJ-Nda-
dl-
If
D-N da = D-N =
aavV
K-v^dl K-v^dl,, I
dv++ uif Uf da , Pf dv
v\
"c r
r
| B BN N d
av av j- [ M t
dt\ v
JN
ujfdcr) w fdcr}== -
Pfdv dv+ + Pf
"
av --
da +
[uf(vN cfn(vn-N) \Ndl J-Nd*+j[u N-cnn-N)-j-Ljjdl
aa (wfV, + K)-vfi dl
s
a
212
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
where the meanings of n, N, T, I/ S , vn and cn are the same as in the general balance equations (3.1), (3.5) and a = E n V, I = i fl E. In these expressions, D and B are the electric and magnetic induction fields, E and H the electric and magnetic fields, J is the volume current density, p* denotes the volume free charge and finally K and w± are respectively the surface current density and surface free charge on E and Vs denotes the transport velocity of the free charges along E. The local conditions which correspond to the integral laws (8.1) can be derived by resorting to (3.4) and (3.6). We thus have VxE V x £ == - dB ^ , dt '
in i n C - EE, ,
(8.2)
VxH VxH = v+<^, = JJ ++ Pfvpf+ §, VVD D = pPf> f ,
V • B5 == 00 ,,
dpf dp f „ pfv)--= 0, -^-+V-(J ■{J++Pfv) = 0,
v£z-nxlE-BxW} • n x \E - B x W\ == 0, 0, vTi -nx\H E-nxlH
+ DxWl DxW}
[D)-n = 8ui 8uj f
8u!ff
oonn EE --TT, ,
(8.3)
= K ,
uf,
\B\ ■ nn == \B\ ■ [B] = «00),, IL-"JI " n—
v s ■WfVs)-
-2Hcnuf
+
Vs- K- -[pfU + J-•n] == 0 ,
- fC)u T-[nx((V,-C)LJ T-[nx((V +1-*)] K)] ==== 0,00, oon o nnTT,, T-[nx({V,-C) K)] f -+ a-C)u f Uf (OV
(8.3a)
CHAPTER 8. FERROELECTRICITY
213
where the velocity W of the line / = E D if is given by (2.46) in terms of t; and the normal velocity c n of E and C is the velocity of discontinuity line T on E.
8.3
A system of a rigid ferroelectric crystal in the presence of
conductors and quasi-static approximation.
It is well known t h a t Maxwell's equations are covariant under Lorentz transformations. Since our goal is to use them together with the classical balance equations, which are Galilean invariant, we search for the conditions under which Maxwell's equations are also approximately Galilean invariant. F r o m now on we refer our analysis to a particular continuous system C [75]. Let D be a ferroelectric rigid crystal which is at rest in the frame of reference / . Let us suppose t h a t around D there are Sj fixed rigid conductors, Cj,
i = 1 , . . . , Sj, which are connected to a pole of battery, as well as s 2
isolated rigid conductors C , j = l , . . . , s 2 ,
each of them carrying a total
charge Q . Let C the system constituted by D, the s = Sj + s 2 conductors and by the (eventually unbounded) region V among the crystal and conductors, to which we a t t r i b u t e the electromagnetic characteristics of vacuum.
214
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig. 28 The electric and magnetic fields produced by C are discontinuous across the boundaries of each part of C as well as across the domain walls Sb, b = l,...,q,
which in our model are represented as discontinuity surfaces of
electric and polarization fields on D. D is partitioned by these walls in p domains Da, a= 1,...,p. It is evident that the Maxwell equations are valid over all the volumes occupied by C with the exception of the aforesaid surfaces of discontinuity on which the j u m p conditions are held. Since the whole system is at rest in the frame of reference / , the Maxwell equations and j u m p conditions become
V x £ = - f ,
inC-E,
(8.4)
CHAPTER 8. FERROELECTRICITY
215
V x E = J + dD ~J+ dt ' V-D = pf Pf,
V f■lfl==00, ,
vzE-nx{S-BxW} • n x I E -- £B> x:WJ W] = == (Q, 0,
£ - IT\ , oonn E
(8.5)
fD WDxW]+K)=0, + iV2-nx(lH / E - n x ((Lff[F + D xx W ] + «') = 0 )
[D] wf ,, [ £ ] .- nn == w, [B] ■ nn = 0 , IB] SLO, 8uif
^
+v V s -•(w/VJ( u V V s ) --2Hc 2 t f cs n1 W // s
+T
V sS . tA: f - "[ ptP/ -•n] n ] === 00,, / cCnn,++. /J*
T-lnx((V K)} T-{nx((Vs~C)u 0, f a-C) f + K)] K == G, *VUf-C)w
on o nT ,
(8.6) (8.6)
where, owing to (2.46) and the condition v = 0 , W W
W = r^~c + c^nn.n^f^ =W r'r -^^f^ +
(8.7) (8-7)
In what follows we assume that i) == 00 in CC -: Uu C- ; 0 p. P/ = &
•*
■
>
t := 1
w) C evolves in a neighbourhood of an electric equilibrium configuration and all the variations connected with this evolution are of low frequency. It is possible to prove that under this assumption the nondimensional analysis leads to a drastic simplification of Maxwell's equations.
216
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
Let us denote by I a scaling length and by T a scaling time; moreover, let A be the scaled version of the filled A and e,n,cr are respectively, the dielectric and magnetic permeability and the electric conductivity. If we denote by the same symbol a dimensional and a nondimensional field, it is an easy exercise to verify that (8.4)-(8.6) in nondimensional form can be written as * B L dB , VxE VxE = = -M-kQB, %T at
iinn C C --EE, ,
(8.8)
*
aEaL x€HEL i c-t*E LdE dE nV xB y R= _ ^ —£— JT + % 1T dt di ' B B PL n V V-EE = ?±p., J ! — cE Pi
*PL 5 p / ,v dt + V o-TE*, 0t
V -BB== o, 0, V-
■E
= 0 , 1
* vL-nxlE-£-^BxW} = 0, onE-T, E1 * » "x-nx(lH + ^±ExW}+^K) = 0, B J B
[JB].„ = 4 « / , [-»] • n = 0 , **
+ v s •(w*V.)-2Hcu*+2£v. >,v s )- -2J5fc, i i i , + vs iKf
7^
(8.9)
217
CHAPTER 8. FERROELECTRICITY
Lp * u
IPfn
* E ,aET * JL P
•n] == 0 ,
r-[nx ( 0 V C ) W / + KT 4 f ff)] = 0,
»T.
(8.10)
It is well known that, at equilibrium, the electric and magnetic fields are zero in all the internal points of each conductor C,-, i= l,...,s,
of C. This
implies that, in dynamical conditions the electric fields are produced by temporal variations of magnetic fields and vice versa so that (see (8.8) 1 2 )
£~^B, Therefore, E ~ L2fieE/T2,
B~^fi€E.
E{\ - fieL2/T2)
(8.11)
~ 0; but neL2/T2 ~ 0 and then
E ~ 0. This result and (8.11)2 allow us to conclude that H ~ 0. Finally, by eliminating V JE between (8.8) 3 and (8.8) 5 we derive /> ~ 0. Equations (8.4) become D= E = B= H = J = 0 ,
in C , ; , i = l,...,s.
(8.12)
On the other hand, owing to i) and the condition cn = 0 since the conductors are at rest, we have
nxE=0,
on dC,-, i= l,...,s nx(£T + Ji') = 0 , D-n
= uif ,
J5-n = 0 , ^ + Vs • K - 0 ,
,
(8.13)
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
218
where n is the external unit normal to the conductor boundaries <9C, and the fields are evaluated on dCi coming from the empty space V or from the crystal D. In the empty space V among the conductors and the ferroelectric crystal D we have p, = a = 0 and on dV n 3D it is uij = 0, K = 0. Moreover, in statical conditions, there is an electric field which is due to the charges carried by some conductors and to potentials of the others. Consequently, (8.11)2 is no longer valid. On the contrary, (8.11)2 still holds since the magnetic fields are produced by the temporal variations of electric fields. Then, we have from (8.4)
VxJ5 = 0,
in V,
(8.14)
v x u
~ €° dt ' V-E = Q , V-J3 r = 0 ,
where e0 is the dielectric permeability of the vacuum. Similar considerations, referred to the ferroelectric crystal D, in which B = (ioH,Pf=Ljf=0,K
= 0 ,lead to the following set of differential equations
and jump conditions:
V x £ = 0,
i n £ > - U Sb, b= i
V §& V x J = x a
~ dt ' V ' z> == o , V - f f == 0 ,
(8.15)
CHAPTER 8. FERROELECTRICITY
n x [ £ ] = 0,
219
nx[lf] = 0,
»•[!>] = 0, « x [ £ ] = 0, n-[D] = 0,
on dD ,
(8.16)
n-[H} = 0,
i / r n x [ f + D x l F ] = 0, n.[ir] = 0,
(8.17)
o n S j , 6 = 1,...,?.
By the expression (8.7) of W, condition (8.17)2 becomes
* E • n x [if] + i/ s • (WJDJ - IDJW) = 0 so that, taking into account (8.17) 3 , we have
" E ' ( » x I*} + WJDJ
) = * £ > * ([1T| + [J5] x cnn)) = 0.
Finally, by noting that i/ s is arbitrary and the vector within the parentheses is independent of i/ £ , (8.17)2 can be written as
n x ( [ 5 + Dxc„n]) = 0,
on Sb , b = 1,.. .,n.
(8.18)
In the sequel we will be interested in the jump of Poynting's vector flux Ex H. First of all, from (8.13) and (8.16) we derive the continuity of E and H across dC:, i = l , . . . , s , and dD; therefore, it follows that
n-[ExH]
=0 ,
on
Q 8Ct \JdD. i =
l
(8.19)
220
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
On the other hand, (8.17)j shows that the tangential components of E are continuous across any surface S^ so that we can write
.[£xf n-\ExE\ n
f l = n-\E n - f f£xB\ , x i f ] ,=r ln-E . £ t1x{H\= x [ i r l = - I E, [B\. ? , -■n nx x[IT].
From this equality and (8.18), (8.17) 3 we obtain
n -\E x H\ = = {DSlDj.E \ -Escns,cn, on Sb6 ,, 6 == 1,...,. 1,. ..,q . n.lExH}
(8.20) (8.:
We note that in a ferroelectric crystal the vectors D and E are related by
D = e0E + P0p,
(8.21)
where P0 is a constant, which expresses the polarization intensity and p is a unit vector (see section 1). The constitutive equation (8.21) allows us to write (8.20) as follows
n n-lExHj •[#x
s. • « . • ff] = cC nP^0lpj.E 0 [P]'
(8.22)
It is very easy to verify that (8.12)-(8.22) are invariant under a Galilean transformation if E, D, w*, K are invariant and H1 = H + « x D, where « is the uniform velocity of the frame of reference / ' with respect to / .
CHAPTER 8.
8.4
221
FERROELECTRICITY
The balance of energy and the entropy inequality.
W e suppose t h a t the following energy balance equation is valid for the whole ferroelectric crystal D :
e dv +
dt\
E d ■)=
D
n(T-p-ExH-h)
dc
(8.23)
3D
where £ = (J Sb \J dD, e and E are respectively the energies per unit volume and surface and r-p
is an extra-flux of energy due to the polarization [75].
The other symbols have the usual meaning. In addition to (8.23) we assume the following entropy inequality for the whole crystal:
dt
I
s dv +
S da)>
h-N
-
da
(8.24)
3D
D
where 9 is the absolute temperature, s and S are respectively the entropies per unit volume and surface. In order to find the local implications of (8.22) and (8.23), we begin by noting t h a t V-(ExH) E- VxH V ■(E xH) = E •V xE-H-VxE-E-VxH 2 2 ! ■0EoE -±e -E-P 0p. - -E
= - -E-D= - ED
P0P-
(8.25)
By the Gauss theorem (3.3) as well as by (8.22), (8.25), (2.33) we obtain
—
aD
V-(ExH)
n- Ex H da = I1
n-\ExH\da
dvI
222
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY • 1 td\t0E•2dv 2 = ^£ + l^le20E2}cnd
+ \c}-E J-E + \E-P JE.P 0pnP0lP 0p-jcnP0lPa sd tr.sd
D
This relation m a y be employed to transform (8.23) to the equivalent form [76]
A( j-( dt\
edv+ e dv + E
D
n- ( r -- p-h)
E da\ do\ =
da +
dD
•
P E-pdv P0E Q- p dv
D
2 + \l^0E2-P0P-E^l±e s}d
(8.27) (8.27)
E where we have introduced
ee = = e-±e e - l €0E0 2£ 2 . .
(8.28)
By comparing (8.27) and (8.24) with the general balance law (7.23) and the equivalent local conditions (7.24) we have the following system of partial differential
equations and j u m p conditions
eE = P P00E-p E - p+ + (V(VT)-p ■ T) ■+p tr(T®Vp)-V-h, + tr(r Vp) - V • hin, D ,,
s ++
09
-*-,y»>8. e2
6 b
2
2 -§ -§ -- 2HEc 2HEcn n --l [e e ++^10eE , ]s}c 0E S S ••pp c „n 0 E - - PP 0E
-n-lT-m-h}
(8.29)
= E ,
-K-
— -2/7SY f-2HSc n-lscn-^nj>S, 6t "
Sb6 ,b .,q , on 5 ,6 = = 1,.. l,...,g
-^]>5,
(8.30) (8.30)
CHAPTER 8. FERROELECTRICITY
-EiE whh-v-ui(h)i{h) -ZiEmmW =-i=-i, t
223
/ l ) -
- 8
0 on Thh,h=l,...,r, , h=l,. onT
'
(8.31)
■,r ,
-Z,si{h)i{h) ww <-9, -EiS -v
Eda Eda+Ylh +£h
f
g8rf/ dl = - 0 ,,
(8.32)
r
v.
? d/ == 00.. *dl
Sda+J2 S da+J2hh s
rh
By eliminating V ■ h and [/»] • n between (8.29) and (8.30)]^ 2 and supposing [0] — 0 across any discontinuity surface, we are led to the inequalities -(■j> - (V» + s6) s9) + + PoQE-p E ■ p++ (V • r) • p(V-T)-p+tr(T®V + tr(r Vp) P-)-!±?l>0, ^ ^ > 0, 22 (6V + S^) + 22H*c f QE -(^ - -PQE. # * cnn ++ lTP 0Es.ps}cnPSlCn [V> ++±±e o £ -P
+4!)+
++ [lT-p} T ■P]
where iP == e-0s s-8s , , ip
tfV== .E-
$* == EE-6S - BS ..
(8.32a) (8.32a)
0 , ++$ >$>(>,
224
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
8.5 Constitutive equations and thermodynamical restrictions.
Since in every Weiss domain Da the polarization has a constant modulus P0
which is independent of electric field (saturation), p and E are
independent variables. This implies that we have to specify constitutive equations relating ip, s, T to the polarization field p if we wish to have a complete description of the system we are considering. In full agreement with what is usually assumed, we propose here constitutive equations of the following kind (8.33)
A = A{p,Vp,0), where A denotes any of the quantities ip, s, r.
In agreement with the well-
known dissipation principle, the inequality (8.32a)j has to be regarded as a restriction on the constitutive equations. To apply this principle, we need to evaluate %j> by employing the constitutive equation (8.33), to substitute the resulting expression into (8.32a)j and to impose that the inequality (8.32a)j is satisfied in every process verifying the condition P-P=l.
(8.34)
We note that (8.34) implies also the following relations which we use in the sequel p - p = 0,
V p - p = 0, V P - p + V p - p = 0.
Thus we first obtain from (8.31)
(8.35)
CHAPTER 8. FERROELECTRICITY
225
• dip dtp i i,-. ~ de
dip • , <¥ T7 Vp dp
since the spatial and temporal coordinates are independent when v = 0. By substituting this expression into (8.32a)j and taking into account the constraints (8.34), (8.35) by Lagrangian multipliers A,i/,/*, we attain the inequality - ( § £ + s)6»+(P ») » + ( F00£7+V-r £ + V . r - g - A-\pp - pH-Vp) . V p ) . jp.
+{T*
Wi-a*.)*^-
_h-V9>Q 9 -
(8.36)
which, in agreement with the dissipation principle, has to be satisfied in every process, i.e. for all the values of p,p,Vp,Vp,9,V9.
From (8.36) we
obtain dip 09 '
s= d^ T
>j
P0E,
d
~ P;j
dip f dip
= dpi-{apii
+
(8.37)
; + f*jPi, \
,t
WiJ'i+Xp*
h/ i ■ve V0<
,
+ ,t p
iU
'
226
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
p E
+p +p "°°^wA£t^> -=wA£r,)-' °Xp°"Xpwhere we have put P0\
= - u
( (8.38) 8>
"
,- + A. The relation (8.38) shows t h a t it is
always possible to assume ft = 0 to evaluate E; we attain the same conclusion for r , which appears in the integral balance law through the term r - p . This term, owing to the constraint (8.35)j, is independent of p. On the other hand, it is straightforward to verify by a simple calculation that (8.32a)j is independent of p. Therefore, (8.37) 2 becomes
Uj-
av
(8.39)
~dP',j'
In order to make explicit the consequences of the surface inequality (8.37) 2 , we choose for \t, S and $ constitutive equations of the following form
tf == V(p * ( p ±±,n,n,0), 9 = t6)t where p
(8.40)
are the values of p on both sides of a domain wall. Differentiating
(8.40) we obtain
SV
<9tf St ' ~ dpd+P+
S +
P St st
. <9# ' dp~' dP-
sPSt st
, <9tf Sn.dV S6 dn 8n St ++ 89 St
+ +
where, according to (2.19) Sp - =
st -p
P ++ cCn Vp. nn n. Vp.
Taking into account these last relations as well as the constraint (8.35) 1 , we
CHAPTER 8. FERROELECTRICITY
227
have the inequality
\d9+^) ♦(■
■T +
aP+
, a* 6n I^+i
+
8n
-A + P + >
st +,
<5p + -r «5« - ( -
a* a P - •+ A p "
-
)
■
6p~ St
e0E2 + ^o * V P S - n - r «'A;,fcnfc + 2fftf] c n + * > 0
which implies for arbitrary processes
5=-%|,
(8.41)
"•r+--=7^ +A+p 3p" • j y
T»' T
a* aP-
-A
p
,
ff •ft+ ^ + 5 e° ^ + P°E*' p* - n « r 'i p i, A"* + 2^*1 c„ + * > where the Thomas derivative of n, with the help of (2.16), (2.21), (1.23), can be written as
%=h-V,-Vsn=h-Vs-a°®n,a=-a°%cnja.
(8.42)
Here ^
is the metric tensor in the curvilinear coordinates (ua) on E and a ,
a =1,2,
is the holonomic basis associated with these coordinates. The
formula (8.42) allows us to write (8.41)4 as follows
228
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
-(Uacn);a +
U + ^e0E2
+ P0ES ■ p3 - nirijPJ!
hnh
+ 2H 0 .
(8.43)
where Ua is the surface vector which is tangent to Sb and has components given by Ua = ^-aa\.
(8.44)
Finally, from (8.31) we have
Ei*Hh)"nhywh>(s-9*) = *'
(8-45)
8.6 Structure of the Weiss domains.
We now assume that the polarization field P = PQp(x) is a constant vector in each of the Weiss domains [74]. This implies that the polarization gradient vanishes throughout the union of all Weiss domains. If p is constant then the field equations in the Weiss domains and the relevant jump conditions on their walls become simply (see (8.38), (8.15)a 3 , (8.17)3) -^- + P(jVip + PoXp = 0 ,
in Da, a = l,2,.,.,m
A^ = 0 ,
(8.46)
229
CHAPTER 8. FERROELECTRICITY
[f0Vv> - P0pj ■ n = 0 ,
onSfc, 6 = 1 , 2 , . . . , , ,
(8.47)
where E = - Vtp. We remark here that the Lagrange multiplier A(x) is in general a different function for each domain. We can obtain V
- V ^ = A(z)P + 7 | - g . The vector dip/dp
(8.48)
is also a constant vector determined by p. If we take the
divergence of (8.48) and note (8.46)2 we see immediately that p-VA = 0 .
(8.49)
On the other hand the integrability conditions of (8.48) are obviously
A, i Pj = A, j p - .
By multiplying this relation by p^ and recalling that p is a unit vector we find that
\j = Ki
PiPj-0
due to (8.49). Hence the Lagrangian multiplier can only be a constant in each of the Weiss domains and the electric potential in a Weiss domain can now be expressed as
230
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
^=-[KPH+^^t)}-
(8-50)
* + K-
Hence the admissible electric field can only be a constant vector in any Weiss domain. However the direction of this field is in general different from that of the polarization field depending on the class of anisotropy of the crystal. Let us next consider the relation (8.47) on one of the walls between adjacent Weiss domains. In view of (8.48) this can now be written as
[(co\ + Po)p + eo±^]-n
= 0,
onSb,b
= l,2,...,q.
(8.51)
If we define the constant vector 6 on Sb by
* = I M + ^ o ) P + '0 7 5 ^ 1
(8-52)
6 n = 0.
(8.53)
then (8.51) yields
This relation implies that the unit normal vector to a wall separating two Weiss domains should remain perpendicular to a constant vector determined by the polarization vectors and Lagrangian multipliers in those domains. If the equation of the domain wall is given by
/(■) = 0 ,
(8.54)
then (8.53) requires that / has to satisfy the following first order linear partial differential equation
CHAPTER 8.
FERROELECTRICITY
231
6-V/ = 0,
(8.55)
whose general solution can easily be expressed as
/(*) = hix3 ~ hxi ~ bi9(t>2xi ~ h\H) = ° '
( 8 - 56 )
where g is an arbitrary function of its argument. The factor b1 is introduced for convenience. It is quite straightforward to see that (8.56) represents a ruled surface whose generating lines are parallel to the vector 6. The principal curvatures of this surface are easily found to be
/c,=0,
K2
= |6|V'[l + ( ^ + V ) 2 + 6 i S ' 2 r 3 / 2 ,
(8-57)
where the primes denote differentiation with respect to the argument. We would now like to impose the continuity of the electric potential cp given by (8.50) across the domain wall. To simplify the notation let us introduce new constant vectors by
c = Ap + - J - g ,
(8.58)
in each Weiss domain and denote the values of these constants in two Weiss domains adjacent to the domain wall S^ under consideration by c"1" and c~. Hence the continuity of
cx
t.
\sb + b+ = «f*i Is + b~ 1 b
b
232
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
or employing (8.56) \ xl + c2 x2 + c 3 + [ 1^1 + 9(hXl - -blX2)] + b + c + Xl + c2+ x2 + c + [ fot + g(b2x1 - bxx2)] + b +
c
b3 = H x\ + C2 x2 + c 3 [ b~xl + 9(b2xl ' -6jX 2 )] + 6 = c f x1 + c2 x2 + c3" [ fti + g(b2xx - bxx2)\ + b~
from which we deduce that - c 3 + )9(hxi - - Mz)= tci+ - c i + r ^ 3 + ~ c3 ^ ( 3~ - c 3 + )9(b2x1 - bxx2)= [Clc+ + -c{-c + X^ (c6 3++ - b c 3 - )]xt
(c3 c
+ ( 2 - 2~ ) 2 +
+ (c2+ -c2)
x2+b+
~ ~■
-b~.
(8.59) (8.59)
This relation requires that g can only be a linear function of its argument such as
g = A{b2xx - 6jx 2 ) + B,
(8.60)
where vl and B are constants. This in turn implies through (8.56) that the admissible domain wall can only be planes. Introducing (4.15) into (4.14) we further obtain that
h+ ~ H + f {4 - c3~) = - Ab2(c+ - c3~) , c + - c 2 = Atj (c + - c 3 ) ,
6
+_6-
=
-B(C3+-c3-)l
(8.61)
CHAPTER 8. FERROELECTRICITY
233
which impose certain restrictions on the position of the domain wall. To conclude we would like to point out under which conditions the polarization and electric fields become parallel in a Weiss domain. It follows from (8.46) that this is possible if and only if
- V ^ = AP + - i - ^ = p p ,
mDa
(8.62)
which implies that ^ g
= (/i-A)P.
(8.63)
If we define a generally nonlinear operator -4. by
l H pp-0d0di P
=
., N
A(p) p
-
(8.64)
then (4.23) becomes [A(p)-(H-\)I]-P [A(p)-(n-X)I]-p
= 0.0.
(8.65)
In general this is not an eigenvalue problem in the true sense of the word. Only in the case of a crystal in which ip is a quadratic function of p then (8.65) is an eigenvalue problem and we can explore all the results provided by the linear algebra and determine p as an eigenvector of the matrix A which does not depend on p now. However for nonlinear crystals (8.65) is a set of nonlinear algebraic equations which may have a solution or not. But just to fix the idea we shall keep on calling p as a generalized eigenvector of A- ■ if (8.65) has a solution. Of course the existence of such an eigenvector
234
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
depends heavily on the crystal class.
8.7 T h e boundary value problem for the Weiss domains.
Let us suppose that i) the whole crystal D is made up of Weiss domains Da, a = 1,...,p, which are separated by the domains walls Sb, b = 1,...,g; ii) the constitutive equations of the localization residuals \t are given by * = A(p±,n)cn;
(8.66)
Hi) the effects of temperature are negligible. In these hypotheses we can conclude that the complete set of partial differential equations, jump and boundary conditions for the system C of ferroelectric crystal and conductors can be written as follows
Aip = 0 ,
V
(p =
in V ,
ondVnCj, on dVnC{
(8.67)
j = l,...,s2,
(8.68)
, { = !,...,st ,
„(c) , 1 dtp Va= - [ V ( 0 ) + 7 5 - - ^{ y ] - * + & a . "o dp '
i n £ > a , a = l,...,p
(8.69)
CHAPTER 8. FERROELECTRICITY
235
Mfc = 0,
onSb,b=l,...,q
(8.70)
r 6E -b + 2i c0E2 + P0ES ■Ps¥ nib L St ' l(eoA
+
z
= o,
Po)P + e o ± ^ t } . n = 0,
U + ±t0E2 + P0Ef.pJb
- ^ ■ n - ( M +P
o
)
P
+
( o
+
Ab = 0,
i | ] . n = 0,
ondD.
We note that equations (8.41)2 3 , in which we have now r regarded as restrictions on the constitutive equation ^ ( p
(8.71)
= 0 , should be
, n). The equations
(8.70) constitute a system of Aq equations in the 4p + g - l unknowns AQ, b , pa,
(cn)b ( we recall t h a t p is unitary and the electric potential is defined to
within
a n arbitrary
constant
so that
one of the A's can be assigned
arbitrarily). Moreover, (8.67)-(8.68), together with (8.71), lead to a boundary value problem for the electrostatic field in the vacuum. Since in general 4g > 4 p + g—1, t h e system (8.70) could not admit a solution, i.e., it is n o t possible t o determine t h e evolution of the system C in which the crystal is made u p by domains with plane interfaces. In this case, we should admit the presence of regions along the boundary 3D in which p and E are functions of x and t (see [74] for a detailed discussion relative to the equilibrium).
236
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
8.8 Weiss' domains in the absence of electric field at equilibrium.
Now we would like to treat the case in which i) the crystal D is at equilibrium;
ii)
there
is no externally
applied
electric
field;
Hi) the
polarization field in the ferroelectric crystal is such that electric field induced by the distribution of polarization vectors vanishes everywhere. This requires in view of (8.46)j t h a t [74]
1 drP =
FPadp~S
A p "-Xp, '
™v 1,2,.,■;P a. a, a= ™D a=l,2,...,p,
,
(8.72)
or [J.(p)-p + XI]-p [J.{p)-p XI]-p = 0 iinD n £ a>, a , aa == ll ,, 22 ,, .. .. .;P , p ,,
(8.73) [8.73)
which implies that the polarization vector in each Weiss domain can only be a generalized eigenvector of the operator A and the Lagrange multiplier in t h a t domain should be equal to a generalized eigenvalue with opposite sign. On the domain walls separating Weiss domains the polarization vectors now have to satisfy the following j u m p conditions
[ p ] -■" n = [p] = 0o ,,
on o n SSbb , ,6b==l , 1,2,. 2 , . . . ■,q, ,g,
(8.74)
This relation implies immediately that domain walls can only be ruled surfaces (see section 8.5). Since now we cannot use the continuity of the electric potential across domain walls it is not of course possible to prove rigorously that domain walls can only be planar surfaces. However if we note
CHAPTER 8. FERROELECTRICITY
237
that the walls keep their planar form even if a very weak electric field is present we shall assume henceforth they remain plane in the absence of the electric field. The jump condition (8.70)4 is now valid for all wall domains and it is simply reduced to
[i/> + A]fc = 0 But we have to remark that
on
Sb,
b = 1 , 2 , . ..,.
(8.75)
another problem now arises. The boundary
condition (8.71), in view of (8.46) written for E = 0, now requires that pn = 0,
on 3D ,
(8.76)
namely the polarization vector should be in the tangent plane of the boundary of the ferroelectric crystal. If the boundary surface of the crystal is not made up by the union of planar surfaces which can carry an admissible polarization vector it is obvious that internal Weiss domains cannot be extended up to the boundary. Therefore for an arbitrary shape of crystal we are compelled to consider a thin layer adjacent to the boundary of the crystal in which the polarization vector cannot be assumed anymore to be piecewise uniform. We expect that the solution for the polarization vector in this layer Dr approaches constant states which are prevalent in the interior parts of the crystal at places sufficiently far from the boundary 3D. Therefore we assume that there is a piecewise uniform polarization field which is again the solution of (8.65) in the neighbourhoods of the Weiss domain which are in the proximity of the boundary. However the polarization field must vary very
238
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
rapidly from
these constant states to a field which is tangent
to
the
boundary. Hence in a very thin boundary layer adjacent to the boundary we have to assume that polarization gradients are very large but in the rest of Dj^ it approaches asymptotically to a piecewise constant state. In this layer we thus have to consider the general equations
% - * ■ $ - , + ' » = • •
(8.77)
V p = 0 .
By properly non-dimensionalizing the above equations we expect to transform eqs. (8.77)j into those having a small parameter in front of the second order derivatives of p. Hence we are faced with a singular perturbation problem. We would not delve into a detailed analysis of this problem in this work. It is sufficient
to remark here that the approximate boundary layer equations
which might be obtained by the usual singular perturbation technique should be solved by satisfying the boundary conditions
p-n = 0,
n - - ^ - = 0 , on <9D
(8.78)
and the resulting equations should be matched with the constant states which satisfy the following j u m p conditions on the wall domains
[pJ-"=0,
CHAPTER 8. FERROELECTRICITY
239
by employing again the usual matching techniques. It is now in order to remark that such an approach would produce several admissible configurations for Weiss domains. The only way to obtain the actual configuration is to resort to the expression of total energy of the crystal a n d to determine which one of the admissible configurations renders this function a n absolute m i n i m u m . T o illustrate the general approach presented in the foregoing analysis we would like to treat a very simple case. We consider a uniaxial crystal such as B a T i 0 3 (barium titanate) [73] whose axis corresponds to that of easiest polarization. If we identify this axis as the z-axis of the preferred coordinate axes of the crystal, its free energy is given in the first approximation by the following relation
tf = \ { « [ ( V P J 2 + ( V p y ) 2 + ( V p z ) 2 ] + /?(p 2 + p 2 ) } .
(8.79)
We further consider a rectangular specimen of such a crystal shown in fig.29. Since we would like to deal with the case in which there exists only a pure polarization field, we have to consider eqs. (8.72) in order to construct Weiss domains in such a ferroelectric crystal. In view of (8.79) we have
Ppx=-P0\Px,
pPy=-P0XPy,
0 = Apz.
(8.80)
We now further assume that py = 0. Then, we have the following admissible directions for the polarization vector
240
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
fig. 29
A = 0,
P , = T 1 , P x = 0 ; X=-0/Po,
Px=Th
P* = 0.
(8.81)
When we take into account the boundary conditions
px = 0 0111 = 0,1, and p 2 = 0 on z = 0, /
(8.82)
we see immediately that the boundary layer vanishes in this case and Weiss domains have the form shown in fig.30. One can easily verify that all jump conditions on the domain walls are now satisfied. Of course the parameter d
CHAPTER 8. FERROELECTRICITY
241
which defines individual Weiss domains cannot be determined immediately. If there are m such domains we can only state that
(8.83)
dm = lx.
We see t h a t we have essentially two different Weiss domains which are labeled by 1 and 2. For piecewise constant fields (8.79) yields
t^ = 0
inD,
and
V= h 0
in D2 .
(8.84)
The surface energies on the domain walls are of course constant and are given by $ = i/ij
on vertical walls,
\P = %l>2 on slanted walls .
(8.85)
In order to determine d we now take into account the total free energy JJ which can be written as
J ^ 1 ) = [ I /? i - 2m + i>2 STd 2m + ^(l
- d)m + ^ l
(m - 1)] l2.
(8.86)
If we use (8.83) to eliminate m in the above expression we find that
ffP) = [ i p I, <* + ^ - ^
+ 2 «
The m i n i m u m of this function corresponds to
2
I t - l M / x + O] ' 2 -
(8-87)
242
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
4
P
~ d2
dx = 2^[27 . ^
' .
(8.88)
Of course we cannot expect that the expression (8.88) for d satisfies (8.83) for an integer value of m. However in common crystals d is very small and m is quite a large number so that (8.83) may be interpreted without making a great mistake as the closest integer number to the ratio l-^/d^. However one can easily see t h a t the configuration in fig.29 is not the unique solution to the problem. The configuration in fig.30 also satisfies all the boundary and j u m p conditions. But in this case the broken lines are not wall domains since there is
n o discontinuity of the polarization vector across them. Hence t h e total
energy does not have the last term in (8.86) and we obtain instead of (8.87)
ff(2) = ( J/?/ a- dT- + ^ J A + 2 > f 2 ^ / 1 - ^ 1 / 1 ) / 2 y 4
n
d
(8.89)
the m i n i m u m of which yields
oL = 2 -\fT.\J=T dr 2
(8-9°)
If we now evaluate the difference of the m i n i m u m values of f? corresponding to two different configurations of Weiss domains we find that
gr(i) _ 3,(2)
=
l/3 1^(42-1)
+ ^l [ h. (^2_ i) _ i] .
(8.9i)
CHAPTER 8. FERROELECTRICITY —
243
Since /j/c^ is a large number we deduce that (1) grffl ff(2) < SF iF (1) ■•
(8.92)
fig. 30 Therefore the configuration in fig.30 is energetically more favourable. Landau and Lifshitz [72] have obtained the value of the parameter d in magnetic crystals by considering the energy in the internal layer which is represented by the domain wall in our analysis. By repeating their analysis
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
244
for ferroelectric crystals we find that d is given by
d = 2 >|27
4
HJ".
(8.93)
Comparing (8.93) with (8.90) we see at once that the surface free energy density is given by
^ = 2 ^ 9 .
(8.94)
This expression for tp, reflects obviously our original idea t h a t the surface energy density introduces into the analysis the approximate effect of the polarization gradient near the domain walls since its value is determined in a sense by the coefficient a in the polarization energy. Of course we could do our calculations with a different size for each Weiss domain such t h a t d1,d2,...,dm
subjected to the condition dl+d2+
... + dm
= L. But it is a simple exercise to show that the absolute m i n i m u m of *j}f corresponds to the case in which all d- are equal to each other. Finally, we recall that in [77], for the same crystal we have considered in this section, other possible domain distributions are found when not all the angles among the faces are r / 2 .
References
[1]
C. Truesdell, A. C. Toupin, The classical field theories of mechanics, Handbuch der Physik, vol.III/1 (ed. Fliigge), Berlin-Gottingen- Heidel berg, Springer, (1960).
[2]
L. Scriven, "Dynamics of a fluid interface", Chem. Engng. Sci., 12, 98108, (1960).
[3]
R. Aris, Vectors, tensors and the basic equations of fluid mechanics, Prentice-Hall Inc. (1962).
[4]
G. P. Moeckel, "Thermodynamic of an interface", Arch. Rat. Mech. Anal., 57, 255-280, (1974).
[5]
W. Kosinski, Field singularities and and waves analysis in continuum mechanics, Ellis Horwood Limited and PWN- Polish Scientific Publi sher, Warsaw, (1985).
[6]
A. C. Eringen, D. G. B. Edelen, "On nonlocal elasticity", Int. J. Engng. Sci., 10 , 233-248, (1972).
245
246 [7]
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
G. Edelen, N. Laws, "On the thermodynamics of systems with non locality", Arch. Rat. Mech. Anal., 43 , 24-35, (1971).
[8]
G. Edelen, C o n t i n u u m physics, Ed. A. C. Eringen, Academic Press, (1976).
[9]
E. S. Suhubi, "An introduction to nonlocal field theories of continua" Department
of Mechanical Engineering, the University of Calgary,
Alberta, Canada, Report # 390 (March, 1987).
[10] M. Gurtin, "On the two-phase Stefan problem with interfacial energy and entropy", Arch. Rat. Mech. Anal., 96 , 199-241, (1986).
[11] F. Dell'Isola, A. Romano, "On a general balance law for continua with an interface", Ricerche di Matematica, Napoli, vol.XXXV, 325-337, (1986)
[12] J. Gibbs, Collected works, New-York, Longamanus Green and Co, vol.11, (1928).
[13] R. Defay, I. Prigogine, Tension superficielle et adsorption, Editions Desoer, Liege , (1951).
[14] F . Dell'Isola, A. Romano, "On the derivation of thermomechanical balance equations for continuous system with a nonmaterial interface",
247
REFERENCES Int. J. Engng. Sci. , 25 , 1459-1468, (1987).
[15] F. Dell'Isola, A. Romano, "A phenomenological approach to phase transition in classical field theory", Int. J. Engng. Sci., 25, 1469-1475, (1987). [16] A. Romano, E. S. Suhubi, "Balance equations of nonlocal mechanics revisited", Ricerche di Matematica, vol. XXXIX, 333-350, (1990). [17] D. Iannece, "On the phase transitions in fluid mixtures", Meccanica, 19, 182-187, (1984). [18] D. Iannece, "Thermodynamics of continua with adsorbing interface", Int. J. Engng. Sci., 23, 449-459, (1985). [19] M. M. Abbot, H. Van Ness, Thermodynamics, cap. IV, Schaum, McGraw-Hill, (1976). [20] M. Zemanski, Heat and thermodynamics, McGraw-Hill (1943). [21] A.
Pippard,
Elements
of
classical
thermodynamics,
Cambridge
University Press (1964). [22] A. Romano, "Properties of The Gibbs potential and the equilibrium of a liquid with its vapor", Rend. Sem. Matem. Univ. di Padova, 69, 257-
248
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
265, (1983).
[23] J. Serrin, "Phase transitions and interfacial layers for Van der Waals fluids", in Recent Methods of Nonlinear Analysis and Applications, SAFA IV, Napoli (1980).
[24] D. Korteweg, "Sur la forme que permet les equations du movement des fluids si l'on tien compte des forces capillares consees par variations de densite considerable mains continues et sur la theorie de la capillarite, ...", Arch. Neder. Sci. Ex. Nat., 6 , 2-24, (1901).
[25] D. Iannece, A. Romano, G. Starita, " T h e Gibbs principle for
the
equilibrium of systems with two simple or composed phases", Meccanica, 25, 3-11, (1990).
[26] C. Truesdell, W. Noll, The non-linear field theories of mechanics, vol.HI/3, Springer-Verlag, (1965).
[27]
R. Esposito, A. Romano, "Gibbs variational principle for the equi librium of continuous systems with an interface", ZAMP, 35, (1984).
[28] J. Rodrigues, "Sobra un modelo variacional para a solidificacao dum lingote
em
varamento
continuo
Prepubl. (MAF 5) 83, Lisbon (1983).
e
sua
approximacao
numerica",
REFERENCES
249
[29] L. Faria, J. Rodrigues, "Aspects of variational approach to a continuous casting problem", in Bassavit, Damlaminn Fremont, Free boundary problems: applications and the theory, vol. Ill, (1985). [30] M. S. Carslaw, J. C. Jaeger, Conduction of heat in solids, Oxford University Press, (1959). [31] P. L. Chambre, "On the dynamics of phase growth ", Quart. J. Mech. and Applied Math., 9, 224-233, (1956). [32] A. Romano, G. Starita, "Approximate systems describing solid-liquid and liquid-vapor state changes" Arch. Mech., 41, 909-924, (1989). [33] F. Kreith , Principi di trasmisione del calore, Liguori Editore, (1975). [34] C.Trusdell, Rational thermodynamics: a course of lectures on selected topics, McGraw-Hill, New York, (1969). [35] M. E. Gurtin , A. S. Vargas, "On the classical theory of nonreacting fluid mixtures", Arch. Rat. Mech. Anal., 43, 179-197, (1971). [36] M. E. Gurtin, P. Villaggio, "On stability in classical linear theory of mixtures", Ann. Mat. Pura Appl., Ill, 57-67, (1976). [37] S. J. Spector, "On Gibbs stability in the classical theory of fluid
250
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
mixtures", ZAMP, 32, 286-295, (1981). [38] I. Miiller, "Thermodynamics of mixtures of fluids", J. Mech., 14, n.2, 267-303, (1975). [39] F. Dell'Isola, D. lannece, "On phase transitions in classical fluid mixtures with surface adsorption", Int. J. Engng. Sci., 27, n.l, 1069 1078, (1989). [40] D. K. Chattory, K. S. Birdi, Adsorption and the Gibbs surface excess, Plenum Press, New York (1984). [41] I. N. Levine, Physical chemistry, McGraw-Hill, New York, (1978). [42] D. lannece, G. Starita, " On the evaporation of a fluid into a gas" Arch. Mech., 42, 1, 77-85, (1990). [43] G. Wulff, "Zur Frage der Geschwindigkeit des Wachsthums und der Auflosung der Krystallflachen'', Zeitschrift fur Krystallographie und Mineralogie, 34, 499-530, (1901).
[44] A. Romano, E. S. Suhubi, "On Wulffs
law about
equilibrium
configurations of crystals", Int. J. Engng. Sci. , 27, 1, 1135-1142 (1989). [45] L. Landau, Collected papers, Gordon and Breach, Science Publishers,"
REFERENCES
251
T h e equilibrium form of crystals", (1967).
[46]
C. Herring, "Some theorems on the free energies of crystal surfaces", Phys. Rev., 82, 87-93, (1951).
[47]
M.
Drechsler,
J.F.Nicholas,
"On
the
equilibrium
shape
of
cubic
crystals" , J. Phys. Chem. Solids, 28, 2609-2627, (1974).
[48] J. F . Taylor, " Crystalline variational problems" , Bull. Amer. Math., 84, 568-588, (1976).
[49]
J. E. Taylor, "Constructing crystalline minimal surfaces", Annals of M a t h . Studies 105, Seminar on minimal submanifolds, E. Bompieri, 201-288, (1983).
[50] T .
Alts, K. Hutter, " C o n t i n u u m
description of the dynamics
and
thermodynamics of phase boundaries between ice and water", J. NonEquilibrium Thermodyn., 13,221-329, (1988).
[51] T . Alts, K. Hutter, "Towards a theory of temperate glaciers. Dynamics and thermodynamics of phase boundaries between ice and water", Mitt. No. 82 der Versuchsanstalt fur Wasserbau, Hydrologie und Glaziologie, (D. Vischer, Hrsg.) ETH Zurich, 1986.
[52] M. E. Gurtin, Evolving phase boundaries in the presence of deformation
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
252
and surface stress", in On the evolution of phase boundaries, ed. M. E. Gurtin and G. B. McFadden, Springer-Verlag, (1992). [53] L. D. Landau, E. M. Lifsits, Fisica statistics, Editori Riuniti, (1978). [54] D. G. Edelen, N. Laws, "Nonlocal Continuum Mechanics", Arch. Rat. Mech. Anal., 43, 36-44, (1971). [55] D. G. B. Edelen, Continuum physics, (ed. by A. C. Eringen) Vol.IV, Polar and nonlocal theories , Academic Press , New York, (1976). [56]
E. S. Suhubi, An introduction to nonlocal field theories of continua, Dep. of Mechanical Engineering, The University of Calgary , Alberta, Canada, Report No 390 (March, 1987).
[57] A. Romano, E. S. Suhubi, "Balance equations of nonlocal Mechanics revisited" Ricerche di Matem., vol. XXXIX, fasc. 2 , 333-350, (1990). [58] W. K. Burton, N. Cabrera, F. C. Frank,"The growth of crystals and the equilibrium structure of their surfaces", Phil. Trans., A243, 299-358, (1951).
[59] D. lannece, A. Romano, "Solidification of small crystals and nonlocal theories", Int. J. Engng. Sci. 28, 535-542, (1990).
253
REFERENCES
[60] D. lannece, A. Romano, "Growth of macroscopic crystals and nonlocal theories", Int. J. Engng. Sci., 29, 1199-1204, (1991). [61] D. lannece, A. Romano, "On the mathematical modelling of crystal growth in a binary nonreacting mixture", Math. Models and Methods in Applied Sciences, 3, 485-496, (1993). [62] W. F. Brown, Magnetoelastic interactions, Springer, Berlin, (1966). [63] W. F. Brown, Micromagnetics, Wiley, New York, (1963). [64] G. Maugin, "Sur la dynamique des milieux deformable magnetises avec spin", J. Mecanique 13,75-96 (1974). [65] G. Maugin, A. C. Eringen, "Deformable magnetically saturated mediaI- Field equations", J. Math. Phys., 13, 143-155, (1972). [66] G. Maugin, A. C. Eringen, "Deformable magnetically saturated mediaII- Constitutive equations", J. Math. Phys., 13, 1333-1347, (1972). [67] A. Romano, "A macroscopic non-linear theory of magnetothermoelastic continua", Arch. Rat. Mech. Anal., 65, 1-24, (1977). [68] W. von Doering,
Micromagnetismus, Handbuch der Physik, Band
XVIII / 2 , Springer, Berlin, (1966).
254
PHASE TRANSITIONS IN CLASSICAL FIELD THEORY
[69] M. Rascle et al., (Eds), "PDE's and phase transitions", (NSF-CNRS Meeting in Nice, France), Lecture Notes in Physics, Springer, Berlin, (1991). [70] L.D. Landau, E. M. Lifshits, "On the theory of dispersion of magnetic permeability in ferromagnetic bodies", Phys. Z. Soviet 8, 153-169, (1935). [71] E. M. Lifshits, "On the magnetic structure of iron", J. Phys., USRR 8, 337-346, (1944). [72] T. Mitsui, I. Tatsuzaki, E. Nakamura, An introduction to the physics of ferroelectricity, Gordon k, Breach, New York, (1976). [73] S. Chikazumi, S. H. Charap, Physics of Magnetism, Wiley, New York, (1964). [74] A. Romano, E. S. Suhubi, "Structure of Weiss domains in ferroelectric crystals", Int. J. Engng. Sci., 30,1715-1729, (1992). [75] A. Romano, "Sull'evoluzione dei domini di Weiss in cristalli ferroelettrici rigidi", Ricerche di Matematica, in press.
[76] G. A. Maugin, Continuum mechanics of electromagnetic solids, NorthHolland, Amsterdam, (1988).
REFERENCES
255
[77] A. R o m a n o , E. S. Suhubi, "Possible configurations for Weiss domains in uniaxial ferroelectric crystals'', Ricerche di Matematica, in press.