You are on page 1of 425

Vibrations of shells and rods

Khanh Chau Le
Second edition (internet version) 2013

To my parents, Le Khanh Can and Truong Thi Tan Nhan

Preface
We live in a world of vibrations and waves, without which there would not
be sound, light, radio, television, communication etc. That is why the study
of vibrations and waves is so important in many branches of physics and mechanics. This book is devoted to the study of small mechanical vibrations of
shells and rods, which are made of elastic or piezoelectric materials. But even
in this very special field there are already many excellent books and monographs written since the monumental work by Rayleigh [47]. The peculiarity
of the present book is that we regard the equations of shells and rods as
two- and one-dimensional approximate equations which can be derived from
the three-dimensional theory by using the variational-asymptotic method.
The latter has been invented especially for those variational problems which
contain small parameters. It turns out that for vibrations of shells and rods
there are many situations in which such small parameters exist. Thus, the
application of the variational-asymptotic method enables one to derive not
only the classical two- and one-dimensional theories of low-frequency vibrations of shells and rods, but also the theories of high-frequency (or thickness)
vibrations.
The present book is organized into ten chapters. After the short introductory chapter containing some historical background we provide preparatory
material on tensor analysis, geometry of curves and surfaces, dynamic theories of elasticity and piezoelectricity, and the variational-asymptotic method
in the second chapter. The rest of the book is divided into two nearly
equal parts which treat the theories of low- and high-frequency vibrations,
respectively. Chapters 3-6 present two- and one-dimensional theories of lowfrequency vibrations and wave propagation in thin bodies, namely elastic
shells and plates (Chapter 3), elastic rods (Chapter 4), piezoelectric shells
and plates (Chapter 5), and piezoelectric rods (Chapter 6). Chapters 7 and
8 deal with high-frequency vibrations of elastic shells, plates, and rods, and
finally, Chapters 9 and 10 study high-frequency vibrations of piezoelectric
shells, plates, and rods, respectively. To help a reader become more proficient, each section ends with problems and exercises, of which some can be
3

4
solved effectively by using the Mathematica. Difficult problems are marked
with an asterisk. It is not our aim to give complete references on the subject,
which is very large. We cite rather those papers which are directly related
to the methods used in the book.
This book is intended for engineers who deal with vibrations of shells
and rods in their everyday practice but also wish to understand the subject from the mathematical point of view. Some of the results concerning
high-frequency vibrations of shells and rods may be new for them. The book
can serve as a textbook for graduate students who have completed firstyear courses in mechanics and mathematics. It may also be interesting for
those mathematicians who seek applications of the variational and asymptotic methods in elasticity and piezoelectricity. Only a minimum knowledge
in advanced calculus and continuum mechanics is assumed on the part of a
reader.
I would like to express here my deep gratitude to my teacher Prof. V.L.
Berdichevsky (Detroit), who has had a great influence on my development of
the subject. Substantial parts of Chapters 3 and 4 are based on his lectures
and publications. I thank Prof. H. Stumpf (Bochum) for his warm hospitality during the writing of the book, and Professors R.J. Knops (Edinburgh),
A.G. Maugin (Paris), W. Pietraszkiewicz (Gdansk), L. Truskinovsky (Minneapolis), D. Weichert (Aachen) and many other friends and colleagues for
their comments and useful discussions. The competent language assistance
by Mrs. Anne Gale (Springer Verlag) is also gratefully acknowledged.
Last, but not least, thanks are due to my wife and my daughter, without
whose patience and love this book would not have appeared at all.
Bochum, May 1999
K.C. Le

Contents
1 Introduction

2 Preliminaries
2.1 Tensor analysis . . . . . . . . . .
2.2 Geometry of curves and surfaces .
2.3 Dynamic theory of elasticity . . .
2.4 Dynamic theory of piezoelectricity
2.5 Variational-asymptotic method .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Low-frequency vibrations

57

3 Elastic shells
3.1 Two-dimensional equations . . . . . . .
3.2 Asymptotic analysis . . . . . . . . . .
3.3 Dispersion of waves in plates . . . . . .
3.4 Frequency spectra of circular plates . .
3.5 Dispersion of waves in cylindrical shells
3.6 Frequency spectra of cylindrical shells .
3.7 Frequency spectra of spherical shells .
4 Elastic rods
4.1 One-dimensional equations
4.2 Asymptotic analysis . . .
4.3 Cross section problems . .
4.4 Dispersion of waves . . . .
4.5 Frequency spectra . . . . .

.
.
.
.
.

.
.
.
.
.

17
17
24
30
37
45

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

59
59
68
78
89
96
109
118

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

123
. 123
. 130
. 140
. 147
. 155

5 Piezoelectric shells
163
5.1 Two-dimensional equations . . . . . . . . . . . . . . . . . . . . 163
5.2 Asymptotic analysis . . . . . . . . . . . . . . . . . . . . . . . 172
5

6
5.3
5.4
5.5

Error estimation and comparison . . . . . . . . . . . . . . . . 183


Frequency spectra of circular plates . . . . . . . . . . . . . . . 193
Frequency spectra of cylindrical shells . . . . . . . . . . . . . . 200

6 Piezoelectric rods
6.1 One-dimensional equations
6.2 Asymptotic analysis . . .
6.3 Cross section problems . .
6.4 Frequency spectra . . . . .
6.5 Longitudinal impact . . .

II

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

High-frequency vibrations

209
. 209
. 215
. 224
. 234
. 240

249

7 Elastic shells
7.1 Two-dimensional equations . . . . . . .
7.2 Long-wave asymptotic analysis . . . . .
7.3 Short-wave extrapolation . . . . . . . .
7.4 Dispersion of waves in plates . . . . . .
7.5 Frequency spectra of plates . . . . . .
7.6 Dispersion of waves in cylindrical shells
7.7 Frequency spectra of cylindrical shells .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

251
251
257
267
275
285
296
305

8 Elastic rods
8.1 One-dimensional equations . . .
8.2 Long-wave asymptotic analysis .
8.3 Short-wave extrapolation . . . .
8.4 Cross section problems . . . . .
8.5 Dispersion of waves . . . . . . .
8.6 Frequency spectra . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

311
311
316
322
329
336
340

.
.
.
.
.

349
. 349
. 351
. 360
. 369
. 374

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

9 Piezoelectric shells
9.1 Two-dimensional equations . . . . . .
9.2 Long-wave asymptotic analysis . . . .
9.3 Short-wave extrapolation . . . . . . .
9.4 Frequency spectra of circular plates .
9.5 Frequency spectra of cylindrical shells

.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

10 Piezoelectric rods
381
10.1 One-dimensional equations . . . . . . . . . . . . . . . . . . . . 381
10.2 Long-wave asymptotic analysis . . . . . . . . . . . . . . . . . . 383
10.3 Short-wave extrapolation . . . . . . . . . . . . . . . . . . . . . 390

7
10.4 Cross section problems . . . . . . . . . . . . . . . . . . . . . . 395
10.5 Frequency spectra . . . . . . . . . . . . . . . . . . . . . . . . . 400
A Material constants
405
A.1 Elastic isotropic materials . . . . . . . . . . . . . . . . . . . . 405
A.2 Piezoelectric crystals . . . . . . . . . . . . . . . . . . . . . . . 406
A.3 Piezoceramic materials . . . . . . . . . . . . . . . . . . . . . . 408
B List of notations

409

Bibliography

413

Index

419

Chapter 1
Introduction
The study of vibrations of elastic shells and rods began with the pioneering
works of Daniel Bernoulli and Euler. They derived the one-dimensional differential equations of the flexural vibrations of beams by what we now call
the variational principle of stationary action. They determined the eigenfunctions and the eigenfrequencies of a beam in the six cases of boundary
conditions corresponding to the free, clamped or fixed edges. The BernoulliEuler theory preceded the exact three-dimensional linear elasticity discovered
by Navier, Cauchy and Lame. Immediately after this great discovery Poisson applied three-dimensional elasticity to the derivation of one-dimensional
equations of vibrations of thin rods. Regarding the rod as a circular cylinder of small cross section, he expanded all the quantities in powers of the
distance from the central line of the cylinder. When terms above a certain
order (the fourth power of the radius) are neglected, the equations for flexural vibrations turn out to be identical with those of Bernoulli-Euler. The
equation for the longitudinal vibrations was derived by Navier; that for the
torsional vibrations was first obtained by Poisson. Saint-Venant proposed
the semi-inverse method for solving the problems of torsion and flexure of
beams within 3-D elasticity. Although his method is not directly related to
the dynamics, its influence on the development of shell and rod theories cannot be overlooked. Concerning the bending of beams, Saint-Venant adopted
two assumptions: i) extensions and contractions of the longitudinal fibres
are proportional to their distances from the plane drawn through the central line at right angles to the plane of bending, and ii) there is no normal
traction across any plane parallel to the central line. The application of the
theories rests upon a principle introduced by Saint-Venant and bearing his
name, according to which statically equivalent tractions applied to the end
of the bent beam or twisted bar produce the same stresses far from their
9

10

CHAPTER 1. INTRODUCTION

end.1 Kirchhoff generalized Saint-Venants ideas for rods undergoing large


displacements. He deduced an approximate expression of the strain in an
element of the rod, and then found the one-dimensional energy functional.
He obtained the equations of equilibrium or motion by varying the energy
functional. From the three-dimensional elasticity, the problem of wave propagation in an infinite cylinder of the circular cross section was treated by
Pochhammer and Chree, who obtained the dispersion relation for the axisymmetric longitudinal waves.
The correct two-dimensional equation of small flexural vibrations of plates
was first proposed by Sophie Germain. Poisson and Cauchy both considered
this problem from the point of view of 3-D elasticity, supposing that all the
quantities which occur can be expanded in powers of the distance from the
middle surface. Poisson again derived Sophie Germains equation. Much
controversy has arisen concerning Poissons boundary conditions. These said
that the resultant forces and moments applied at the edge of the plate must
be equal to the forces and moments arising from the strain. Kirchhoff was the
first to show that these conditions are too numerous and cannot in general be
satisfied. He proposed another method of derivation of the two-dimensional
equations for bent plates, which has remained in use until today. Similar to
the rod theory, Kirchhoffs reasoning was based on two assumptions: i) the
straight fibres of a plate which are perpendicular to the middle surface before
deformation remain so after deformation, and ii) all the elements of the middle surface remain unstretched. These assumptions enabled him to propose
an Ansatz for the displacement and strain fields in the plate. Substituting these fields into the action functional and integrating over the thickness,
Kirchhoff found the expression of the two-dimensional functional of the bent
plate in terms of the curvatures produced in its middle surface. The equations of motion and boundary conditions were then derived by the variational
principle, and they were applied to the problem of flexural vibrations of a circular plate. The results were compared with those obtained experimentally
by Chladni. Much later Rayleigh and Lamb found the dispersion equation
for waves in an infinite plate according to 3-D elasticity.
The two methods of construction of the approximate theory of plates
described above found their natural generalization in the approximate twodimensional theory of shells, the foundation of which was laid down in the
book by Love [35].2 Inconsistencies in Loves equations were noted and corrected by several Russian scientists in the 1940s (see e.g. [17, 42]). The sim1

The mathematical proof of this principle was given much later by Toupin [59] (see
also [4, 15]).
2
It also contains the most important references on the subject prior to its appearance.

11
plest and most elegant version of the shell theory was proposed by Koiter [23]
(see also [52]). He later gave the error estimate of the first-order shell theory
in the energetic norm based on Prager-Synges identity [24]. From the exact
3-D elasticity the dispersion equation for waves in circular cylindrical shells
of infinite length was derived and studied in detail by Gazis [2, 16].
Tracing the history of the development of the approximate shell and
rod theories, one can easily observe the competition between two methods,
namely asymptotic and variational ones. In order to use the variational
method, one needs an a priori Ansatz for the displacement and strain fields
as functions of the transverse co-ordinates. Substituting this Ansatz into the
3-D action functional, one can derive the equations determining the dependency of the displacement field upon the longitudinal co-ordinates by varying
the functional obtained after the averaging procedure. This method resembles the direct method of Rayleigh and Ritz and the semi-discrete method of
Kantorovich-Krylov [26] of solving approximately variational problems. The
disadvantage of the variational method is the necessity of having an Ansatz
for the displacements, while simplicity and brevity are its advantages. By
the asymptotic method we mean the expansion of the displacements into an
asymptotic series, the substitution of this series into the equations of the 3-D
theory, and the subsequent asymptotic derivation of the recurrent system of
equations for the corresponding terms of the series. The asymptotic method
needs no a priori assumptions; however, it is very cumbersome. The synthesis of these two methods, called the variational-asymptotic method and first
proposed by Berdichevsky [5, 6], seems to avoid the disadvantages of both
methods described above and proved to be very effective in constructing
approximate equations for shells and rods. The variational-asymptotic procedure provided in [5,6] enables one to construct not only the first-order shell
and rod theories, but also the correct refined theories which are asymptotically exact in the long-wave range. This method later found wide application,
also in other problems of mechanics and physics (see e.g. [6]).
With the variational-asymptotic method the theories of vibrations of
shells and rods have been put upon a firm foundation. One can show that the
classical two-dimensional equations of motion of elastic plates and shells can
be used to describe their vibrations in the low-frequency long-wave range.
The exact solutions of the three-dimensional equations of elasticity for infinite plates, by Rayleigh and Lamb, or for infinite cylindrical shells, by Gazis,
confirm this conclusion. A similar situation exists in regard to the classical
one-dimensional equations of motion of elastic rods as compared with the
exact 3-D elasticity studied by Pochhammer and Chree. However, numerical
analysis of Rayleigh-Lambs and Pochhammer-Chrees dispersion equations
shows that, as the frequency increases, many new branches of the disper-

12

CHAPTER 1. INTRODUCTION

sion curves arise (see [41]). These branches are connected to each other in
the complex wave-number plane, signifying the complicated interaction between waves of different branches near the free edge of the plate, shell or
rod. As the wave number and the frequency increase, the velocities in the
three-dimensional theory have upper limits for all branches, in contrast to
the classical two-dimensional theory. Hence the latter cannot be expected to
give good results for the frequencies of modes of vibration of high order.
Timoshenko [57, 58] was the first to include the effect of transverse shear
deformation and rotatory inertia to derive a one-dimensional theory of flexural motions of bars which gives more satisfactory results for short waves
and high modes of vibrations. But Timoshenkos theory and its generalizations for plates and shells (Reissners and Berdichevskys refined shell theories [5,48]) have the shortcoming that they cannot describe satisfactorily the
cut-off frequency (corresponding to the zero wave number) and the long-wave
asymptotes of the first branch of thickness vibrations.
It was Mindlin [41], who succeeded in deriving two-dimensional equations
of motions of plates which give satisfactory results for dispersion curves of
both low-frequency and thickness branches. In his pioneering papers the
following method of derivation has been proposed. The displacements are
expressed by the expansions in the series of Legendre polynomials of the
thickness co-ordinate. These series expressions are then substituted into the
three-dimensional action functional followed by an integration over the thickness and a truncation to produce the required order of approximation. Since
Legendre polynomials are not appropriate eigenfunctions of the branches of
thickness vibrations, the two-dimensional theory obtained cannot describe
cut-off frequencies and long-wave asymptotes of those branches. The correction coefficients are introduced to improve the match between the frequency spectra of an infinite plate as obtained from the approximate and
exact equations.
Although Mindlins theories have been successfully applied in many engineering problems (see [41] and quotations therein), his introduction of the
correction coefficients remains a little mystic. Berdichevsky was the first to
show that the long-wave asymptotic analysis can be applied for branches of
high-frequency thickness vibrations of elastic plates near the cut-off frequencies [6]. Based on the variational-asymptotic method he found the distributions of the displacements and derived the equations of high-frequency longwave vibrations for all thickness branches. This method is then applied for
elastic rods [25] and elastic shells [7]. The later checking, by Kaplunov [21,22],
confirms the results for plates, but displays some arithmetic mistakes in the
calculation of the coefficients for the equations of shells, the correction of
which leads to the full agreement of the results.

13
The equations derived in [6, 7] are asymptotically exact and describe correctly the behaviour of plates and shells in the long-wave range near the cutoff frequencies. However, these same equations without modification yield an
unsatisfactory description of the dispersion curves and the group velocities in
the short-wave range. At the same time, the formulation of boundary-value
problems is associated with the behaviour of the corresponding differential
operators at short wavelengths. Thus, even asymptotically exact equations
in the long-wave range may lead to the ill-posed boundary-value problems [5].
Therefore the construction of the theory of shells and rods involves not only
the derivation of equations in the long-wave range, but also another logically
independent step the extrapolation of those equations to short waves.
It is possible to carry out either trivial extrapolations, when the system of equations derived for long waves is applied to short waves without
any changes, or non-trivial extrapolations, when terms that are small in the
long-wave range but appreciable for short waves are introduced (removed).
For shells in the short-wave range it is impossible to describe the threedimensional stress state exactly by the two-dimensional theory, and only a
qualitative agreement can at best be expected. For this reason, different
two-dimensional equations are allowed in the theory of shells. However, it
is natural to demand an asymptotic equivalence in the long-wave range of
different short-wave extrapolations.
In [8, 33] the best hyperbolic short-wave extrapolation is proposed for
the equations derived in [7]. This involves the classical branches and several
thickness branches of vibrations and takes into account their cross-terms at
short waves. The structure of the equations is similar to that of Mindlin for
plates, but in contrast to his theory, the asymptotic accuracy is achieved in
the long-wave range by the asymptotic analysis and not by the introduction of
correction coefficients based on ad hoc assumptions. This brings additional
advantages: i) those problems, for which the exact dispersion equations are
not known or available (for instance, shell vibrations) can also be analyzed,
ii) the asymptotically exact 3-D stress and strain state can be restored from
the 2-D integral characteristics. The application of the 2-D theory to various
problems, such as the dispersion of waves or the frequency spectra of plates
or shells, shows that it enables one not only to predict the asymptotically
exact distributions of the stress and displacement fields in the long-wave
range, but also to describe qualitatively correctly their behaviour in the shortwave range. For instance, in the problem of edge resonance in semiinfinite
plates which admits the localized waves of length comparable with the plate
width, the 2-D theory predicts the frequency of edge resonance lower than
that obtained from 3-D elasticity by about 1% [29]. The construction of the
approximate theory by the variational-asymptotic method is then generalized

14

CHAPTER 1. INTRODUCTION

for elastic rods [31] and sandwich plates [51].


The second group of problems that can effectively be solved by the variational-asymptotic method comprises vibrations of piezoelectric shells, plates
and rods. Piezoelectric crystals and ceramics are materials whose behaviour
clearly demonstrates the coupling between mechanical and electric fields
[37, 38]. Therefore piezoelectric shells, plates and rods are widely used in
acoustics as generators, filters or detectors of vibrations [36]. The piezoelectric effect was discovered in 1880 by brothers P. and J. Curie. The early development of piezoelectricity and its applications up to about 1940 were summarized in [11]. The book by Mason [36] remains an important reference to
many piezoelectric materials and their use in ultrasonics (see also [20]). Variational principles of the linear piezoelectricity were first formulated in [56]
(see also [28, 32, 41]). The two-dimensional equations of longitudinal lowfrequency vibrations of piezoelectric plates were obtained in [36]. The asymptotically exact two-dimensional equations of piezoceramic shells have been
derived in [50] with the help of Goldenveizers asymptotic method [17]. One
can also find in [50] the generalization of Saint-Venants principle to piezoelectric shells and rods (see also [3]). In [28,32] the two-dimensional equations
of statics and low-frequency vibrations for piezoelectric shells in the general
case of anisotropy were derived based on the variational-asymptotic method.
The generalization of the well-known Prager-Synges identity to linear piezoelectricity was found. With the help of this identity the error estimate of the
two-dimensional theory of piezoelectric shells was established in [32]. In this
book the one-dimensional equations of low-frequency vibrations of piezoelectric rods will also be presented.
High-frequency thickness vibrations of piezoelectric plates and shells are
widely used in ultrasonics [9]. Because of the complexity of 3-D boundaryvalue problems, the analysis of these thickness vibrations are normally carried
out on a one-dimensional basis under the assumptions that the plate is laterally infinite in extent with all points in any given plane parallel to the faces
moving with equal displacements and with no phase differences from point to
point [9]. This is not the case in real plates, since the latter are not infinite in
extent and the additional boundary conditions at their free edges should also
be satisfied. Due to the interaction between different branches of vibrations
at the free edge one can observe interesting effects such as the edge resonance
in a circular disk [53]. Mindlin [41] and Tiersten [56] were the first to derive
the approximate two-dimensional equations of high-frequency vibrations for
quarzt crystal plates. The method of derivation is the same as for elastic
plates. Again, the correction factors have to be introduced to improve the
match between the frequency spectra of an infinite plate as obtained from the
approximate and exact equations. In this book we present the full derivation

15
of the two-dimensional equations of high-frequency vibrations of piezoelectric plates and shells based on the variational-asymptotic method [27, 30].
We also derive the one-dimensional equations of high-frequency vibrations of
piezoelectric rods. The applications of the theory in the problems of resonant and antiresonant frequency spectra of plates, shells and rods show that
the 2- and 1-D approximate equations of vibrations are asymptotically exact
in the long-wave range and yield qualitatively good agreement even in the
short-wave range.

16

CHAPTER 1. INTRODUCTION

Chapter 2
Preliminaries
2.1

Tensor analysis

Euclidean point space and translation space. In this book we deal


with vector and tensor fields on domains of the three-dimensional Euclidean
point space E. Elements of E, called spatial points, are denoted x, y, z . . ..
In a chosen fixed cartesian co-ordinate system a point z corresponds to a
triple (z 1 , z 2 , z 3 ), with z i being its i-th co-ordinate. The translation space of

Figure 2.1: Cartesian co-ordinate system.


E is denoted by V; it is a three-dimensional vector space. Elements of V are
called (spatial) vectors and are denoted with boldface letters like u, v, w, . . ..
The scalar and vector products of two vectors u, v V are denoted u v and
u v, respectively. Referring to the cartesian co-ordinate system there is a
17

18

CHAPTER 2. PRELIMINARIES

one-to-one correspondence between any point z and its position vector


z = z 1 i1 + z 2 i2 + z 3 i3 = z i ii ,
where ii (i = 1, 2, 3) are the standard basis vectors (Figure 2.1). Unless otherwise specified we always use the Einstein summation convention: summation
on repeated indices is understood.
Vector fields. Now we can define a vector field u on a domain U E as a
map
u : U 7 V.
This means that in every point z U there exists a vector u(z) V (see
Figure 2.2).

Figure 2.2: A two-dimensional vector field.


Examples of such vector fields are displacement field, velocity field, acceleration field etc. In the cartesian co-ordinate system specified above we can
refer a vector field u(z) to the basis {ii } as follows
u(z) = ui (z 1 , z 2 , z 3 )ii .

(2.1)

Functions ui (z 1 , z 2 , z 3 ) are called cartesian components of the vector field


u(z).
Curvilinear co-ordinates. The fact that we are dealing only with vector
and tensor fields on domains of the Euclidean space does not prevent us from
using co-ordinate systems other than cartesian ones.1 We now introduce a
general curvilinear co-ordinate system {xa } as a one-to-one smooth map that
maps an open set U E to R3 according to
(z 1 , z 2 , z 3 ) 7 (x1 (z i ), x2 (z i ), x3 (z i )).
1

As we shall see soon, this is advantageous in the theory of shells and rods.

2.1. TENSOR ANALYSIS

19

The inverse functions of xa (z i ) are denoted z i (xa ). The following convention


for indices is adopted: while the indices a, b, . . . , h are used when referring
to the curvilinear co-ordinates {xa }, the remaining indices i, j, . . . , z will be
used when referring to the cartesian ones {z i }. Co-ordinate lines x1 are the
curves whose components in the cartesian co-ordinates are z i (t, x2 , x3 ), where
t is the variable and x2 and x3 are fixed (Figure 2.3). Similar definitions hold
for x2 and x3 . Thus, every spatial point can be considered as the intersection
of three co-ordinate lines x1 , x2 , x3 . The tangents to these curves ea can be
chosen as the basis vectors of the curvilinear co-ordinate system.

Figure 2.3: A curvilinear co-ordinate system.


Taking the partial derivatives of the position vector with respect to xa we
obtain
z i
i
ea =
ii = z,a
ii ,
(2.2)
xa
where the comma preceding indices denotes the partial derivatives with respect to the corresponding co-ordinates. Note that ea V and are functions of x1 , x2 , x3 . Because of one-to-one correspondence, the Jacobian of the
transformation z i 7 xa (z i ) is not vanishing,2 so {ea } is the basis for each
i
(x1 , x2 , x3 ). Since xa,i and z,a
are inverse matrices, from (2.2) we have
ii = xa,i ea .
2

(2.3)

Moreover, we require that the Jacobian is positive, so that the orientation of the basis
vectors remains unchanged.

20

CHAPTER 2. PRELIMINARIES

Now we can refer u to the basis {ea }


i
u = ua ea = ua z,a
ii .

(2.4)

Comparing this with equation (2.1) we have


i a
ui = z,a
u .

(2.5)

This is the transformation rule for components of a vector when changing


from cartesian to curvilinear co-ordinate systems.
Tensors. In order to introduce tensors, let us first define the dual space V
as the space of all linear maps that map V into R (elements of V are called
co-vectors or one-forms). We can choose a dual basis {a } for V such that
a (eb ) = ba ,
where ba stands for a Kronecker delta (ba = 1 if a = b, ba = 0 otherwise).
Now we define the tensor of type (p, q) as a multilinear map
t : V V V V 7 R
(p copies of V and q copies of V). We use boldface letters to denote vectors
and tensors. The components of t referred to the basis {ea } and its dual
{a } are defined by3
a ...a

tb11...bqp = t(a1 , . . . , ap , eb1 , . . . , ebq ).


We say that t is contravariant of rank p and covariant of rank q. We regard
scalars as (0,0)-tensors. The tensor t can also be written as
a ...a

t = tb11...bqp ea1 . . . eap b1 . . . bq ,


where denotes the tensor product. Addition and scalar multiplication of
tensors are defined in the obvious way. The set of all (p, q)-tensors forms a
tensor space denoted by Tqp (V). A (p, q)-tensor field on the domain U E is
defined as a map t : U 7 Tqp (V).
Because V is the inner product space, for each co-vector V there
exists uniquely an associated vector t V such that
(u) = t u,
3

for all u V.

(2.6)

One can see from this definition that, while tensors do not depend on the choice of
the basis vectors and covectors, their components do.

2.1. TENSOR ANALYSIS

21

To show this, let us introduce the reciprocal basis vectors ea V such that
ea eb = ba .
Now for = a a we choose t = a ea . It is easy to show, then, that
equation (2.6) is satisfied. Due to this one-to-one correspondence we shall
identify co-vectors with their associated vectors and the dual basis {a } with
the reciprocal one {ea }. Let us denote by gab and g ab the following quantities
gab = ea eb ,

g ab = ea eb .

Note that
g = gab ea eb
forms a symmetric positive definite second-rank tensor called the metric tensor. It is easy to see that
ea = g ab eb ,

ea = gab eb .

Given a tensor we can calculate an associated tensor by applying operations of


raising or lowering of indices with the help of g ab or gab . Consider for example
the (0,1)-tensor (co-vector) = a a . The raise of the index leads to the
new vector ] = g ab b ea . By virtue of the last identity this is nothing but
the vector t = a ea , which has been identified with . Thus, the operations
of raising and lowering of indices with the help of metrics does not change
tensors at all. Therefore we are allowed to use the same letter for a tensor
and its associated tensor obtained by raising or lowering of indices. It is also
clear that
g = gab ea eb = g ab ea eb = ba ea eb = 1,
so that the metric tensor is equal to the identity tensor. It is interesting
to note that the components of the metric tensor referred to the cartesian
basis {ii } are simply the Kronecker delta ij , and that the reciprocal basis
coincides with {ii }. Therefore the raising or lowering of indices of tensors
referred to the cartesian basis does not change their components at all.
Given two tensors t and s, we can obtain a new tensor by multiplying
them and contracting the p-th contravariant index of t with the q-th covariant
index of s. Such an operation is called one index contraction of t and s and
the result is denoted by t s. When two indices are contracted, the result
will be denoted by t : s. In the subsequent chapters we shall mainly use the
index notation, according to which contracted indices are simply repeated.4
Contraction of the p-th contravariant index of t with the q-th covariant index
4

One also speaks of them as dummy indices.

22

CHAPTER 2. PRELIMINARIES

of the same tensor is defined similarly. For example, the trace of a second
rank tensor t is the scalar obtained by
trt = taa .
Gradient of tensors. Let us consider a tensor field t : U 7 Tqp (V) defined
on an open subset U of E. The tensor field t is said to be of class C 1 if there
p
is a tensor field t : U 7 Tq+1
(V) such that
t(x + v) = t(x) + (t(x))v + (x, v),
where

x U, v V,

1
(x, v) = 0
|v|0 |v|
lim

holds for all x U. The tensor field t, if it exists, is uniquely determined


by t and is called the gradient of t. This definition does not depend on
the co-ordinate systems. Let us work out the gradient for a vector field u.
Referring to the cartesian co-ordinates {z i } we have
u(z) = (ui (z)ii ) = ui (z),j ii ij .
This is due to the fact that the basis vectors ii are constant everywhere.
Consider now a curvilinear co-ordinate system {xa }. According to (2.4) we
have
u(x) = (ua (x)ea ) = ea ua (x) + ua (x)ea .
This equation holds true due to the product rule for the gradient. For its
first term
ua (x) = ua (x),b eb .
To transform the second term we use (2.2) and (2.3) to get
i
i
i
ea = (z,a
ii ) = z,ab
ii eb = z,ab
xc,i ec eb .

Introducing Christoffel symbols


i
cab = z,ab
xc,i ,

and combining the last two equations, we obtain


(ua ea ) = (ua,b + uc acb )ea eb = ua;b ea eb .
The co-ordinate expression for the covariant derivative of tensor fields may be
worked out in a similar manner. Notice that we use a semicolon preceding an

2.1. TENSOR ANALYSIS

23

index to denote the co-ordinate expression for the gradient. We also denote
by ua;b the following expression
ua;b = g bc ua;c .
The reader can find many useful techniques of the tensor calculus in [40].
Gauss theorem. Given a smooth tensor field t, we can obtain a new tensor
field by contracting the last contravariant and covariant indices of t. This
operation is called the divergence of t and the result is denoted by divt.
As an example let us consider the divergence of a second rank tensor field
t = tab ea eb
divt = tab
;b ea .
What we obtained is a vector field. Let U be a domain in the Euclidean space
E with a regular boundary U, on which a smooth tensor field t is defined.
Then Gauss theorem states
Z
Z
divt dv =
t n da,
(2.7)
U

where dv and da are the volume and area elements in E, respectively, and n
is the unit outward normal vector to U. The proof of this theorem may be
found in any standard texbook on geometry (e.g. [54, 55]).
Problems
1. Find the covariant components gab and the contravariant components
g ab of the metric tensor referred to the
i) spherical co-ordinates defined by
z 1 = x1 sin x2 cos x3 ,

z 2 = x1 sin x2 sin x3 ,

z 3 = x1 cos x2 ,

ii) cylindrical co-ordinates defined by


z 1 = x1 cos x2 ,

z 2 = x1 sin x2 ,

z 3 = x3 .

2. Show that the Christoffel symbols can be calculated by the formula


1
abc = g ad (gbd,c + gcd,b gbc,d ).
2
3. Prove that the divergence of the vector field u = ua ea is given by
1
divu = ( gua ),a ,
g
where g = det gab .
4. Show that the gradient of the co-vector field t = ta ea is given by
(ta ea ) = (ta,b tc cab )ea eb = ta;b ea eb .

24

2.2

CHAPTER 2. PRELIMINARIES

Geometry of curves and surfaces

Moving triad of a curve. Consider a curve c(x) in the Euclidean space E


given by a vector equation of the type
z = r(x).
We assume that r(x) is a sufficiently smooth vector-value function of x.
Without limiting generality we can also assume that x is the arc-length of
the curve. The unit tangent vector t to the curve is then calculated as:
t = r0 (x),
where the prime denotes differentiation with respect to x. Differentiating
the identity t t = 1, we see that the vector t0 is orthogonal to the tangent
vector t. Assume that the vector t0 does not vanish. We then choose the
principal normal n to the curve as the unit vector co-directional with t0 . The
proportional factor is denoted and called the curvature of the curve at the
point in question
t0 = n.
(2.8)
The quantity 1/ is called the radius of curvature. Now we choose the vector
of binormal b as the unit vector orthogonal both to t and to n so that the
triad t,n,b is positive oriented (Figure 2.4). These three vectors form a
moving triad of the curve as x changes.

Figure 2.4: A moving triad.


Since n is a unit vector we prove, in exactly the same manner as for t,
that n0 is orthogonal to n. Differentiating the identity t n = 0 with respect
to x we have
t0 n + tn0 = 0,
that is,
tn0 = t0 n = n n = .
Expressing n0 as the linear combination of t and b and recalling the last
equation we have
n0 = $b t.
(2.9)

2.2. GEOMETRY OF CURVES AND SURFACES

25

The quantity $ is called the torsion of the curve. Its inverse 1/$ is called
the radius of torsion.
In the same manner we can see that b0 is orthogonal to b. Differentiating
the identities t b = 0 and n b = 0, we have
tb0 = t0 b = 0,
and
nb0 = n0 b = ($b t)b = $.
Therefore we conclude that
b0 = $n.

(2.10)

Equations (2.8), (2.9) and (2.10) are called Frenet formulae expressing the
change of the moving triad of the curve.
Surface co-ordinates. Consider a surface S in the Euclidean space given
by a vector equation of the type
z = r(x1 , x2 ),

(2.11)

or, in components
z i = ri (x ),

i = 1, 2, 3;

= 1, 2.

Figure 2.5: A triad t1 , t2 and n on a surface.


We shall use Latin indices running from 1 to 3 to refer to the spatial
co-ordinates and the Greek indices running from 1 to 2 to refer to the surface
co-ordinates x1 and x2 . The functions ri are supposed to be sufficiently
smooth. Co-ordinate lines x1 of the surface are the curves r(t, x2 ), where t
is the variable and x2 is fixed. A similar definition holds for x2 . Thus, every

26

CHAPTER 2. PRELIMINARIES

point of the surface can be considered as the intersection of two co-ordinate


lines x1 and x2 . The tangents to these curves t can be chosen as the basis
vectors of the two-dimensional curvilinear co-ordinate system of the surface.
From (2.11) we have
t = r, .
The surface is assumed to be regular, so the vectors t1 and t2 are linearly
independent. We choose the unit normal vector n to the surface as the unit
vector co-directional with t1 t2 (Figure 2.5)
n=

t1 t2
.
|t1 t2 |

Due to this choice, the triad t1 ,t2 and n is positive oriented.


The first fundamental form of a surface. From equation (2.11) follows
dz = r, dx = t dx .
The square of the element of distance on the surface is therefore given by
ds2 = dzdz = t t dx dx = a dx dx ,

(2.12)

a = t t

(2.13)

where
are called components of the metric tensor of the surface. It is clear that the
form (2.12) is symmetric and positive definite. It is easy to check that the
determinant of the matrix a is equal to
det a = |t1 |2 |t2 |2 sin2 = |t1 t2 |2 ,
where is the angle between t1 and t2 . Therefore the area element of the
surface is given by
p
da = det a dx1 dx2 ,
where denotes the wedge product of two differential forms.
The second fundamental form of a surface. Let us consider the vector
value one-form
dn = n, dx .
We define the second fundamental form of the surface as the scalar product
of dz and dn taken with a minus sign
dzdn = t n, dx dx = b dx dx .

(2.14)

2.2. GEOMETRY OF CURVES AND SURFACES

27

The symmetry of b follows at once from the identities t n = 0. Indeed


b = t n, = t, n = r, n = b .
The normal curvature of the surface in the direction dz is given by
=

dzdn
b dx dx
=
.
ds2
a dx dx

The principal curvatures are roots of the quadratic equation


det(b a ) = 0.
The two roots 1 , 2 of this equation correspond to the maximum and minimum values of . The directions which give the principal curvatures are called
the principal directions. Except so-called umbilic points, where 1 = 2 ,
there exist in each point of the surface two principal directions, which are
mutually orthogonal. A curve on the surface which is tangent to one of the
principal directions in each point is called a line of principal curvature. The
quantities
1
1
H = (1 + 2 ) = a b ,
2
2

K = 1 2 =

det b
det a

are called the mean and the Gaussian curvature of the surface, respectively.
Intrinsic geometry of a surface. Let us introduce the tangent space in
each point of the surface. It has the structure of the two-dimensional vector
space. The basis vectors of this space are denoted = . In this tangent
space we introduce surface vectors and tensors in a similar manner as in the
three-dimensional case. In particular, the metric tensor is defined by
a = a ,
where are the dual basis vectors. Using the metrics a and its inverse
a we can define the operations of raising and lowering of indices for surface
vectors and tensors. We also define an operation of covariant derivative
such that its associated parallel translation of vectors along a curve on the
surface preserves lengths and angles between them. Let u = u be a vector
field on the surface, then it can be shown that

u ) ,
u = u; = (u, +

where
are the Christoffel symbols of the surface connection

= a (a, + a, a, ).
2

28

CHAPTER 2. PRELIMINARIES

We use the semicolon preceding Greek indices to denote the co-ordinate expression for the covariant derivatives on the surface.
In general, the intrinsic geometry of the surface is non-Euclidean. For
example, the second covariant derivatives are not commutative. Let u be a

vector field on the surface. We define the curvature tensor R.


by

u .
u; u; = R.

The contraction R = R.
is called the Ricci curvature. It can be shown
that R = 2K.
Given a vector field u defined on a surface S bounded by a smooth closed
curve S, and suppose that the field u is regular so that divu exists. Then
the surface divergence theorem states
Z
Z
divu da =
u ds,
S

where ds is the length element in E and is the surface unit outward normal
vector to the boundary S. In component form this formula reads
Z
Z

u; da =
u ds.
S

Problems
1. Given a curve
z = r(t),

t [0, a].

Show that the arc-length of the curve can be calculated by


Z tr
dr dr

dt1 .
x(t) =
dt1 dt1
0
2. Find the curvature and torsion of the spiral
z 1 = r cos t,

z 2 = r sin t,

where r and a are constants.


3. Prove Gauss formula for a surface

t, =
t + b n.

z 3 = at,

2.3. DYNAMIC THEORY OF ELASTICITY

29

4. Prove Weingartens formula for a surface


n, = a b r, .
5. The third fundamental form of a surface is defined as
dndn = n, n, dx dx = c dx dx .
Show that
c = a b b = Ka + 2Hb .
6. Show that on a spherical surface of radius R
u; u; =

2.3

1
u .
R2

Dynamic theory of elasticity

Kinematics. Linear elasticity, discovered by Navier, Cauchy and Lame, distinguishes itself by its rich mathematical structures in spite of the simplicity
of its underlying basic principles. In this section we summarize briefly the
main concepts and properties of elastodynamics. Let B E be a domain occupied by a linear elastic body in its stress-free undeformed state. A motion
of the body is defined by a continuously differentiable vector field w(z, t)
called the displacement field that depends on time t as well. The relative
displacements of different material points yield the strain inside the body,
which is described by the following symmetric second-rank tensor field
1
= (w + (w)T ),
2
called the strain field, where the superimposed index T denotes the transpose
of a second-rank tensor. Referring to the cartesian co-ordinate system with
the basis {ii } the components of the strain field are given by
1
ij = (wi,j + wj,i ) = w(i,j) ,
2
where the parentheses in subscripts denote the symmetrization operation.
Referring to the curvilinear co-ordinate system {xa } with the basis {ea } we
have
1 i
ab = (z,a
wi,b + z,bi wi,a ) = w(a;b) .
(2.15)
2

30

CHAPTER 2. PRELIMINARIES

Because the displacement field w depends on t, we introduce the velocity


and acceleration fields characterizing its time rates according to

v = w,

a = w,

where the dot denotes the partial derivative with respect to t.


Balance equations. Let (z) be the mass density of a body (the scalar
field), (z, t) the stress tensor field, b(z, t) the vector field of the body force.
We require the balance equations of momentum and moment of momentum
Z
Z
Z
dv =
w
b dv +
n da,
U
U
U
Z
Z
Z
dv =
z w
z b dv +
z (n) da,
U

to hold for an arbitrary regular subdomain U B. Assuming the regularity


of all the above fields and applying Gauss theorem (2.7) to these equations,
we derive the balance equations in the local form
= b + div,
w

(2.16)

T = .
Thus, according to the last equation the stress field must be symmetric.
Hookes law. The generalized Hookes law states that the stress tensor
at any point of a body is proportional to the strain tensor at the same
point
(z) = C(z):(z).
(2.17)
The fourth-rank tensor field C(z) is called the tensor field of elastic moduli. When C(z) does not depend on z, we call the body homogeneous. If
not otherwise stated, we shall consider in the following homogeneous elastic
bodies.
We further assume that there exists a quadratic form W () such that
equation (2.17) can be presented in the following way
=
where

W
,

1
1
W = : C : = ab C abcd cd
2
2

(2.18)

(2.19)

2.3. DYNAMIC THEORY OF ELASTICITY

31

corresponds to the strain energy (per unit volume) of the elastic body. We
require that
1
ab C abcd cd aab ab ,
2

a>0

for all 6= 0. This inequality can be understood as the necessity of work


being done to deform the body from its stress-free natural state.
In component form equation (2.17) reads
ab = C abcd cd .

(2.20)

Due to the symmetry of and as well as (2.18), the components of the


tensor C of elastic moduli satisfy the following symmetry properties
C abcd = C bacd = C abdc = C cdab .
By this the number of the independent components of C is reduced from
81 to 21. For isotropic elastic bodies we have only two independent elastic
moduli (Lame constants) and such that
C abcd = g ab g cd + (g ac g bd + g ad g bc ).
Depending on particular problems one can use other pairs of constants to
characterize isotropic elastic bodies, for example Youngs modulus E and
Poissons ratio . The relations between them are given in Appendix A1.
Boundary-value problems. Substituting (2.15) and (2.20) into the balance of momentum (2.16) we obtain three differential equations of motion in
terms of displacements
w a = (C abcd wc;d );b + ba .

(2.21)

We assume that the boundary B is decomposed into two subboundaries Sk


and Ss such that
Sk Ss = , Sk Ss = B.
On the part Sk the displacements should be prescribed. In most cases we
assume that
wa = 0 on Sk .
(2.22)
On the remaining part Ss the traction is prescribed
ab nb = C abcd wc;d nb = ta

on Ss .

(2.23)

32

CHAPTER 2. PRELIMINARIES

In most cases we assume that ta = 0. We also have to specify the initial


conditions at t = t0
t=t0 = v0 .
w|t=t0 = w0 , w|
(2.24)
The system of equations (2.21), boundary conditions (2.22), (2.23) and initial
conditions (2.24) composes a well-posed dynamic boundary-value problem
and can be used to determine the displacement field.
When the body force b and the traction t on Ss vanish, the vibration of
the body is called free. The motion can then be represented as superposition
of particular motions of the type
wa (z, t) = w a (z)eit ,
with the frequency of such a mode of motion.5 Substituting this equation
into (2.21) with ba = 0 we reduce it to
(C abcd wc;d );b + 2 w a = 0.

(2.25)

Similarly the boundary conditions become


w a = 0 on Sk ,

(2.26)

C abcd wc;d nb = 0 on Ss .

(2.27)

The system of equations (2.25) and boundary conditions (2.26) and (2.27)
composes a well-posed eigenvalue problem.
Hamiltons variational principle. Let us introduce the action functional
of a linear elastic body as follows:

Z t1 Z
Z
I[w] =
(T W + b w) dv +
t w da dt,
(2.28)
t0

Ss

where T is the kinetic energy density given by


1
w,

T = w
2

(2.29)

W the strain energy density given by (2.19). This action functional is defined
on the space of admissible continuously differentiable displacement fields satisfying the kinematical boundary condition (2.22). We require also that w is
given at t = t0 and t = t1
w|t=t0 = w0 ,
5

w|t=t1 = w1 .

In all equations with complex numbers we should take only their real part.

2.3. DYNAMIC THEORY OF ELASTICITY

33

is
Hamiltons variational principle6 states that the true displacement field w
the stationary point of the action functional (2.28)
I = 0.

(2.30)

We now show that all the equations of linear elastodynamics can be derived
from (2.30). Indeed, calculating the variation of (2.28) we have

Z
Z t1 Z
w
:w + bw) dv +
(w
tw da dt = 0.
I =
B

t0

Ss

The formulae for T and W as well as the symmetry of have been used in
deriving this equation. We transform it further with the help of integration by
parts using Gauss theorem, the constraint (2.22) and the vanishing variations
of w at t0 and t1 to

Z t1 Z
Z
+ div + b)w dv +
(t n)w da dt = 0.
I =
(w
t0

Ss

Due to the arbitrariness of w in B and on Ss , we derive from here the


equations of motion (2.16) and the boundary condition (2.23) on Ss . This
variational principle will be used in the subsequent chapters to derive the
equations of vibrations for elastic shells and rods.
For the eigenvalue problem (2.25), (2.26) and (2.27) of the free vibration
with the frequency the following variational principle holds true: among
w a satisfying the boundary condition (2.26) the actual functions w a make
the functional
Z
1
1
( 2 w a wa C abcd wa;b wc;d ) dv
2
B 2
stationary.
Properties of the eigenvalue problem. Let us denote by L the following
differential operator
(Lu)a = (C abcd uc;d );b ,
and by the eigenvalue 2 . Then the eigenvalue problem (2.25) can be
written in the operator notation as follows
Lu + u = 0,
6

(2.31)

Hamiltons principle requires that the displacements are given at t0 and t1 . Clearly,
this does not lead to the formulation of the initial conditions of the type (2.24). Alternative
variational formulations leading to the initial boundary-value problem can be found in [19].

34

CHAPTER 2. PRELIMINARIES

We have to find numbers and corresponding vector


where u stands for w.
fields u 6= 0 satisfying the boundary conditions (2.26) and (2.27).
We can show immediately that the eigenvalues n are non-negative. Indeed, multiplying (2.31) by u and integrating over B we obtain
hLu, ui + hu, ui = 0,

(2.32)

where hu, vi denotes the scalar product


Z
hu, vi =
u v dv
B

in the vector function space. Integrating the first term of (2.32) by parts and
using the homogeneous boundary conditions (2.26) and (2.27) one can easily
check that
Z
C abcd u(a;b) u(c;d) dv 0.
hLu, ui =
B

From here the non-negativeness of the eigenvalues follows. Those eigenfunctions corresponding to the zero eigenvalue belong to the kernel of the
operator L. Note that for traction-free boundary conditions translations and
small rotations are such eigenfunctions.
Consider two eigenfunctions u1 , u2 corresponding to two different eigenvalues 1 6= 2 . Then u1 , u2 are orthogonal in the following sense
hu1 , u2 i = 0,

hLu1 , u2 i = 0.

To show this, we multiply the equation


Lu1 + 1 u1 = 0
by u2 and use the identity
hLu1 , u2 i = hu1 , Lu2 i.

(2.33)

Taking into account the similar equation for u2 we obtain


1 hu1 , u2 i = 2 hu1 , u2 i.
Since 1 6= 2 , the orthogonality follows. Note that an operator L possessing
the property (2.33) is called self-adjoint.
Under some additional conditions7 one can show that self-adjoint operators have a countable set of eigenvalues
0 1 2 . . . ,
7

lim n = +,

For instance, the body is compact and the operator L is positive definite.

(2.34)

2.3. DYNAMIC THEORY OF ELASTICITY

35

and the corresponding eigenfunctions form a basis in the chosen function


space (see, e.g., [12, 49]). This property, called completeness, can be used to
construct solutions of the initial boundary-value problem. The set (2.34) is
called the spectrum of the operator L.
Plane waves in infinite elastic media. Let us analyze the special case,
for which B = E (an infinite elastic medium), b = 0 (no body force), and
and C are constant everywhere. By a plane wave we mean a displacement
field of the form
w(z, t) = a(k z t),
where a, k are constant fixed vectors, is a positive constant called the
frequency of wave propagation, and is a sufficiently smooth scalar function.
It is easy to see that the displacements are the same for all points lying
on a plane orthogonal to k. This vector describes the direction of wave
propagation. Now we select a, k and so that equation (2.21) holds true.
Notice that
= a 2 00 , wk,lj = ak kl kj 00 .
w
Substituting this into the equation of motion (2.21), eliminating the common
factor 00 and dividing by |k|2 we get
1 ijkl
C j l ak = c2 ai ,

where c = /|k| corresponds to the phase velocity of wave propagation and


= k/|k| is the unit vector. Thus, the equation of motion (2.21) is satisfied
if and only if a is an eigenvector with an eigenvalue c2 of the symmetric
matrix ik = C ijkl j l /. For an arbitrary direction k we expect that all the
eigenvalues of this matrix are real and positive (that means ik is positive
definite). In this case we say that equation (2.21) is of hyperbolic type. We
then have in each direction k three different waves associated with three
eigenvectors a1 , a2 , a3 and three eigenvalues c21 , c22 , c23 .
Problems
1. Derive the balance equation of energy for elastic bodies
Z
Z
Z
d
dv +
da.
(T + W ) dv =
bw
tw
dt B
B
Ss
This reduces to the conservation of energy when b = 0 and t = 0.
2. Using the balance of energy, show that the dynamic boundary-value
problem (2.21)-(2.24) yields a unique solution.

36

CHAPTER 2. PRELIMINARIES
3. Show that the strain energy of an isotropic elastic body is positive
definite if and only if the following inequalities hold true:
2
+ > 0,
3

> 0,

or, in terms of Youngs modulus E = (3 + 2)/( + ) and Poissons


ratio = /2( + ),
E > 0,

1 < < 0.5.

Note that for real isotropic elastic materials 0 < < 0.5.
4. Show that there are two velocities of propagation for plane waves in
infinite isotropic elastic media, given by
s
r
+ 2

cd =
and cs =
,

corresponding to dilatational (a is co-directional with k) and shear (a


is orthogonal to k) waves, respectively, with c2s a double eigenvalue.
5. Prove that the lowest eigenvalue of the problem (2.25), (2.26), and
(2.27) can be calculated by
R abcd
C wa;b wc;d dv
2
B R
= a inf
,
w
(2.26)
w a wa da
B
where w a (2.26) means that the minimum should be sought among
functions satisfying (2.26) (Rayleighs formula).

2.4

Dynamic theory of piezoelectricity

Kinematics. Piezoelectric crystals and ceramics are materials whose behaviour clearly demonstrates the coupling between mechanical and electric
fields. Therefore piezoelectric shells and rods are widely used in acoustics as
generators or detectors of vibrations [36]. We present here basic equations of
the dynamic theory of piezoelectricity.
Let B E be a domain occupied by a linear piezoelectric body in its
stress-free undeformed state. A motion of the body is completely determined
by two fields, namely, the vector field w(z, t) called the displacement field,
and the scalar field (z, t) called the electric potential. We assume that these

2.4. DYNAMIC THEORY OF PIEZOELECTRICITY

37

fields are continuously differentiable. As for elastic bodies, the second-rank


strain tensor field is given by
1
= (w + (w)T ).
2
The electric field E is determined from by the formula
E = .
We also need the velocity and acceleration fields defined by

v = w,

a = w,

where the dot denotes the partial derivative with respect to t. The electric
potential depends on time as well, but its time rates do not enter the
equations of motion of piezoelectric bodies, as we shall see below.
Balance equations. As far as the mechanical balance equations are concerned, we require the balance equations of momentum and moment of momentum
= div,
w

(2.35)

= ,
to hold everywhere inside the domain B (we assume from the beginning that
the body force vanishes).
As we shall be concerned with vibrations of non-conducting piezoelectric
bodies at frequencies far below optical frequencies, the conduction current
and rate of change of magnetic induction may be neglected in the Maxwell
equations [39], so we can reduce the latter to the equation of electrostatics
divD = 0,

(2.36)

where D(z, t) is called the electric induction field.


Constitutive equations. In linear piezoelectricity the following equations
are widely accepted:
= cE : Ee,
D = e : + S E.

(2.37)

They describe the well-known coupling between the mechanical and electric
fields. Here cE is the fourth-rank tensor of elastic constants, e the third-rank

38

CHAPTER 2. PRELIMINARIES

tensor of piezoelectric constants, and S the second-rank tensor of dielectric


constants.8 These same equations in component form read
cab
Ec ,
ab = cabcd
E cd e

Da = eabc bc + ab
S Eb .
We assume that the constitutive equation (2.37) can also be derived from
the so-called electric enthalpy W (, E) according to:
=

W
,

D=

W
,
E

(2.38)

where

1
1
W (, E) = : cE : E e: ES E.
(2.39)
2
2
The components of the tensors cE , e, and S satisfy the following symmetry properties
cabcd
= cbacd
= cabdc
= ccdab
E
E
E
E ,
ecab = ecba ,

ba
ab
S = S .

Therefore in the general case the number of independent components of cE


is 21, of e is 18, and of S is 6. The further reduction in number of independent components for the 20 crystal classes which permit piezoelectricity is
discussed in detail in the books by Mason [36] and Nye [43]. The abbreviated
indicial notation turns out to be more convenient than the extended tensor
notation when discussing symmetry. This abbreviated indicial notation consists of replacing pairs of indices ab or cd by m or n according to the following
prescription:
ab or cd 11
m or n
1

22
2

33
3

23,32
4

31,13
5

12,21
6

Gothic letters are used to denote the abbreviated indices. In Appendix A2 a


graphic presentation of the elastic, piezoelectric and dielectric constants for
the 32 crystal classes is reprinted from [9]. Appendix A3 lists experimental
data for some typical piezoelectric ceramics that are transversely isotropic.
Note that isotropic piezoelectric bodies do not exist, because the third-rank
tensor of piezoelectric constants e would vanish in that case.
8

The subscript E in cE indicates elastic stiffnesses at constant electric field, and the
subscript S in S denotes dielectric permittivities at constant strain.

2.4. DYNAMIC THEORY OF PIEZOELECTRICITY

39

Boundary-value problems. We consider only vibrations of piezoelectric


bodies excited by different values of electric potential on portions of the
boundary coated by electrodes. Concerning the boundary conditions for the
mechanical quantities we assume that the boundary B is decomposed into
two subboundaries Sk and Ss such that
Sk Ss = ,

Sk Ss = B.

On the part Sk the displacements should be prescribed. In most cases we


assume that
wa = 0 on Sk .
(2.40)
On the remaining part Ss the traction is prescribed
ab nb = ta

on Ss .

(2.41)

In most cases we assume that ta = 0.


Consider now the boundary conditions for the electric potential. Let the
(1)
(n)
boundary B be decomposed into n + 1 subboundaries Se , . . . , Se and Sd .9
(1)
(n)
The subboundaries Se , . . . , Se are covered by electrodes. We assume that
the electrodes are infinitely thin so that their kinetic and electroelastic energies can be neglected compared with those of the body. On these electrodes
the electric potential should be prescribed
= (i) (t) on Se(i) , i = 1, . . . , n.

(2.42)

On the uncoated portion Sd of the boundary we require that the surface


charge vanishes
D n = 0 on Sd .
(2.43)
We also have to specify the initial conditions at t = t0
w|t=t0 = w0 ,

t=t0 = v0 .
w|

(2.44)

The system of equations (2.35), (2.36), (2.37), boundary conditions (2.40)(2.43) and initial conditions (2.44) becomes well-posed and can be used to
determine the fields w and . When the piezoelectric constants e vanish, the
mechanical and electrical problems are uncoupled, and the former reduces
simply to the anisotropic elasticity. Note also that the dynamic equations
of piezoelectricity are not of the hyperbolic type. If the piezoelectric constants e vanish, then the equations for w are hyperbolic, while that for
is elliptic. Physically, this means that the velocity of propagation of electric
9

The subscript e indicates electrodes, and d stands for dielectrics.

40

CHAPTER 2. PRELIMINARIES

disturbances is much larger than that of the mechanical nature, so that the
former is set equal to infinity.
Variational principle. The action functional of a piezoelectric body is
defined as follows:
Z t1 Z
(T W ) dv dt,
(2.45)
I[w, ] =
t0

where T is the kinetic energy density given by


1
w,

T = w
2

(2.46)

W the electric enthalpy given by (2.39). This action functional is defined on


the space of admissible continuously differentiable displacement and electric
potential fields satisfying the boundary conditions (2.40),(2.42). We also
require that w is given at t = t0 and t = t1
w|t=t0 = w0 ,

w|t=t1 = w1 .

Now the variational principle of piezoelectricity states that among all the
and electric potential
admissible fields (w, ) the true displacement field w
field make the action functional (2.45) stationary
I = 0.

(2.47)

From this variational principle one can derive all the equations of linear
piezoelectricity in the same manner as was shown in the preceding section
for linear elasticity. The variational principle (2.47) will be used in the subsequent chapters to derive the equations of vibrations of piezoelectric shells
and rods.
It is interesting to note that the electric enthalpy W from (2.39) is not a
positive definite quadratic form. If we apply Legendre transformation [13,62]
to W (, E) with respect to the variable E, we obtain the so-called internal
energy density
U (, D) = max[D E + W (, E)].
E

It turns out that


1
1
U (, D) = : cD : D h : + D S D
2
2
is a positive definite quadratic form with respect to and D. The following
constitutive equations equivalent to (2.38) can be derived from the internal

2.4. DYNAMIC THEORY OF PIEZOELECTRICITY

41

energy density U (, D)
U
= cD : D h,

U
= h : + S D.
E=
D
=

(2.48)

In terms of w and D the following dual variational principle turns out to be


valid: among all functions w and D satisfying (2.40) and the constraints
divD = 0,

D n = 0 on Sd ,

make the func and electric induction field D


the true displacement field w
tional
Z t1 Z
Z t1 X
Z
n
I[w, D] =
(T U ) dv dt
(i) (t)
D n da dt
t0

t0

(i)

i=1

Se

stationary.
Let us introduce two other quadratic forms, G(, E) and F (, D), by
applying the Legendre transformation to U (, D) in all variables and with
respect to , correspondingly
G(, E) = max[: + E D U (, D)],
,D

F (, D) = max[: U (, D)].

(2.49)

The quadratic form G(, E) is positive definite and is given by


1
1
G(, E) = : sE : + E d : + ET E.
2
2

(2.50)

Following [36] we call G(, E) the complementary energy density (or Gibbs
function). Similar to W (, E), the quadratic form F (, D) is not positive
definite and is given by
1
1
F (, D) = : sD : D g : + D T D.
2
2

(2.51)

We call F (, D) the elastic enthalpy.10 Dual variational principles for 3-D


piezoelectricity formulated in terms of G(, E) and F (, D) can be found
in [60].
10

The subscript D in sD indicates elastic compliances at constant electric induction,


and the subscript T in T stands for dielectric impermeabilities at constant stress.

42

CHAPTER 2. PRELIMINARIES

Eigenvalue problems. Now let us analyze motions of the special harmonic


type
it

w(x, t) = w(x)e
,

it
(x, t) = (x)e

The equations of motion then become

;bab + 2 w a = 0,

a = 0,
D
;a

Ea = ;a ,
= cabcd cd ecab Ec ,

ab = w(a;b) ,

ab
E
a

D = eabc bc + ab
S Eb .

(2.52)

We also assume that the values of the electric potential on the electrodes are
harmonic functions of time
(i) (t) = (i) eit

on Se(i) , i = 1, . . . , n.

Consequently the boundary conditions read


w a = 0 on Sk ,

ab nb = 0 on Ss ,
a na = 0 on Sd ,
D
= (i)

(2.53)

on Se(i) , i = 1, . . . , n.

Equations (2.52) and (2.53) compose a boundary-value problem for the determination of w a and .
For this boundary-value problem the following
a make the
variational principle holds true: the actual functions w a and D
functional
Z
Z
n
X
1 2 a
a

a na da
(i)
D
[ w wa U (ab , D )] dv
(i)
2
Se
B
i=1
a satisfying the constraints
stationary among w a satisfying (2.53)1 and D
a
;a
= 0,
D

a na = 0 on Sd .
D

(2.54)

We now define a mechanical eigenfrequency m under short-circuit conditions as an eigenvalue of the problem (2.52),(2.53) with
(i) = 0 on Se(i) , i = 1, . . . , n.

2.4. DYNAMIC THEORY OF PIEZOELECTRICITY

43

If a frequency r of vibration can be found such that the resultant surface


charges become infinite on electrodes
Z
a na da = for i = 1, . . . , n,
D
(2.55)
(i)

Se

it will be called electrical resonant. A frequency a at which the surface


charges vanish is called, in contrast to the previous case, electrical antiresonant. From the definition (2.55) it is easy to see that resonant frequencies
are also mechanical eigenfrequencies under short-circuit conditions, but the
inverse is not always true.
For mechanical eigenvibrations under short-circuit conditions the follow a of a
ing variational principle holds true: the actual functions w a and D
vibration with an eigenfrequency m make the functional
Z
1 2 a
a )] dv
w wa U (ab , D
(2.56)
[ m
B 2
a satisfying (2.54). For anstationary among w a satisfying (2.53)1 and D
tiresonant vibrations we have to modify the variational principle as follows:
a satisfying the constraints (2.54) and
among all w
a satisfying (2.53)1 and D
Z
a na da = 0 for i = 1, . . . , n,
D
(i)

Se

a of a vibration with an antiresonant frequency


the actual functions w a and D
a make the functional (2.56), with m replaced by a , stationary.
We now prove that the eigenfrequencies of vibrations of piezoelectric bodies under short-circuit conditions are non-negative. Indeed, multiplying the
first two equations of (2.52) by w
a and ,
respectively, substracting the second from the first and integrating over B we obtain
Z
2
a Ea ) dv = 0.
(
;bab wa + m
w a wa D
;a
B

Integrating by parts and making use of the homogeneous boundary conditions, we reduce this to
Z
2
a ) + m
[2U (ab , D
w a wa ] dv = 0.
B

a ) is positive definite, the mechanical eigenfrequency cannot


Since U (ab , D
be negative. In exactly the same manner we can show that the antiresonant
frequencies should also be non-negative.
Problems

44

CHAPTER 2. PRELIMINARIES
1. Derive the balance equation of energy for piezoelectric bodies
d
dt

Z
(T + U ) dv =
B

n
X
i=1

Z
(i)

da.
Dn

(i)
S

2. Using the balance of energy, show that the dynamic boundary-value


problem of linear piezoelectricity yields a unique solution.
3. Find the independent components of cE , e, and S for transversely
isotropic piezoelectric materials (e.g. piezoceramics).
4. Establish the relationship between cE , e, S and sE , d, T .
5. Find an orthogonality condition for mechanical vibrations of piezoelectric bodies with different eigenfrequencies.
6. Find a generalization of Rayleighs formula for the lowest mechanical
and antiresonant frequencies of vibrations of piezoelectric bodies.
7. Prove that m a , where m and a are the lowest mechanical and
antiresonant frequencies.

2.5

Variational-asymptotic method

Introduction. Various problems of mechanics and physics contain small


parameters. These may be, among others, the thickness of a shell, the diameter of cross-section of a rod, the characteristic lengthscale of grains in
polycrystals, or physical parameters such as wavelengths and frequencies of
vibrations. Over decades many well-known asymptotic approaches have been
invented to handle such problems. At the same time it is clear that for those
variational problems with small parameters there should be a direct variational method, which is based on asymptotic analysis of the corresponding
functionals and leads to approximate variational problems.
The so-called variational-asymptotic method for analysis of variational
problems with small parameters is elucidated in this section.11 It enables
one to analyze problems of extremizing functions with a finite number of
variables as well as functionals defined on an infinite dimensional function
space which contain small parameters. In all cases, when exact solutions of
variational problems are known or when other asymptotic methods can be
11

This method was first formulated in [5, 6].

2.5. VARIATIONAL-ASYMPTOTIC METHOD

45

applied, the variational-asymptotic method always leads to the full agreement of the results. But it has a number of advantages compared with the
traditional asymptotic analysis of the differential equations. First of all, one
has to proceed with the asymptotic analysis for only one function, namely
the Lagrangian of the functional, instead of a system of differential equations. Thus, neglecting a small term in the Lagrangian means neglecting
terms in several differential equations, which are not always easy to recognize as small ones. Next, the approximate equations obtained by applying
the variational-asymptotic analysis of functionals always have the variational
structure, which is not necessarily the case for other asymptotic methods.
Thus, it is easier to prove the correctness of approximate equations obtained
by the variational-asymptotic method, as well as to modify them (for example
to provide short wave extrapolations) if needed.
Variational-asymptotic procedure. Let I[u, ] be a functional that depends on a small parameter . This functional is defined on a set M in some
function space. Assume that the functional I[u, ] has a stationary point
denoted by u . The latter depends on  as well; that is why we attach the
index  to u. Assume that u approaches an element u0 M in the asymptotic sense when  0. The questions arise: can we construct a functional
whose stationary point is u0 ? How to determine the refined functional, whose
stationary point approximates u with the given accuracy?
It is clear that the answers to these questions depend on the asymptotic
analysis of small terms in the given functional. We investigate first the
functional obtained by neglecting formally all small terms in I[u, ], that is,
the functional I0 [u] = I[u, 0]. Let M0 be a set of stationary points of the
functional I0 [u]. We can then express an arbitrary element u M as a
sum u = u0 + u? , where u0 M0 , u? M? . Fixing u0 and assuming that
u? is small in the asymptotic sense, we keep in I[u0 + u? , ] the principal
terms involving u? and neglect all other terms which are small. We then
obtain a functional I1 [u0 , u? , ]. Now we can define stationary points u? of
the functional I1 [u0 , u? , ]; these depend on u0 .
Assume that u? is uniquely determined by u0 : u? = u? (u0 , ). We continue
the variational-asymptotic procedure by assuming that u = u0 + u? (u0 , ) +
u?? , where u0 M0 , u?? M?? and u?? is smaller than u? in the asymptotic sense. Keeping principal terms involving u?? in the functional, we can
determine u?? in exactly the same manner as before. We can continue this
procedure as long as it is required.
Let us consider the functionals I[u0 , ] and I[u0 +u? (u0 , ), ]; both are defined on M0 . When the difference between their stationary points turns out
to be small in the asymptotic sense, then one can expect that the stationary

46

CHAPTER 2. PRELIMINARIES

point of the functional I[u0 , ] is the first approximation to the stationary


point of the functional I[u, ]. When the difference between the stationary points of the functionals I[u0 , ] and I[u0 + u? (u0 , ), ] is not small,
but the difference between those of the functionals I[u0 + u? (u0 , ), ] and
I[u0 + u? (u0 , ) + u?? (u0 , ), ] turns out to be small, then the stationary point
of the functional I[u0 + u? (u0 , ), ] becomes the first approximation. If this
is not the case, one has to proceed further. After the first approximation
has been found, the next approximation can be determined as was described
above.
Sometimes the procedure described above becomes more complicated due
to the fact that u? is not uniquely determined from u0 and may contain some
additional degrees of freedom: at fixed u0 stationary points u? form a set M1 .
We express an arbitrary point u M in the form: u = u0 + u? + u?? , where
u0 M0 , u? M1 , u?? M?? . One can then determine u?? in exactly the
same manner as before. As a rule, after a finite number of steps one has no
more degrees of freedom left and an asymptotic series u = u0 + u? + u?? + . . .
can be rewritten as u = v + w(v, ) + w? (v, ) + . . ., where v is an element of a
set N , and all other terms w, w? , . . . of the asymptotic series are determined
uniquely from v. When the difference between the stationary points of the
functionals I[v, ] and I[v + w(v, ), ] both are defined on N turns out to
be small in the asymptotic sense, one can expect that the stationary point
of the functional I[v, ] is the first approximation to the stationary point
of the functional I[u, ]. When the difference between the stationary points
of the functionals I[v, ] and I[v + w(v, ), ] is not small, but the difference
between those of the functionals I[v+w(u0 , ), ] and I[v+w(u0 , )+w? (v, ), ]
turns out to be small, the stationary point of the functional I[v + w(u0 , ), ]
becomes the first approximation. If this is not the case, one has to proceed
further.
In some cases it is difficult to recognize at the very beginning the smallness of some terms. If the stationary points of the approximated functional
without these terms can be found, we can calculate the neglected terms by
using these stationary points to check their smallness and to verify our assumption. In order to compare stationary points of functionals, sometimes
we need only to compare values of these functionals on the stationary points:
if their differences are small, we expect that the stationary points are close
to each other in the asymptotic sense. At least for convex functionals it is
not difficult to prove this statement.
Characteristic scale. In order to recognize small terms in the asymptotic
analysis of functionals, we often need to know how functions and fields change
their values and which distance is characteristic for these changes. Let us

2.5. VARIATIONAL-ASYMPTOTIC METHOD

47

define the characteristic scale of change for functions.


Consider a function f (x) defined on an interval [a, b]. We assume that
f (x) is continuously differentiable on this interval. We denote by f the
amplitude of change of f (x) on [a, b]:
f =

max |f (x1 ) f (x2 )|.

x1 ,x2 [a,b]

For a sufficiently small positive real number l the following inequality:



df f

(2.57)
dx
l
should hold true. The best constant l in the inequality (2.57) (the largest
real number l, for which the inequality (2.57) still holds) is called the characteristic scale of change of the function f (x) on the interval [a, b]. This
definition of the characteristic scale leads also to an a priori estimate for the
derivative of the function.
If functions change their values considerably, it is more adequate to introduce the characteristic scales of change locally in each point of the interval.
It is also not difficult to generalize this definition for functions or fields depending on more than one variable.
Examples
Newtons polygon. We first illustrate the variational-asymptotic method in
the finite dimensional case, namely in the problem of finding the asymptotics
for the stationary points of polynomials of the type
X
amn m un ,
(2.58)
f (u, ) =
m,n

as  0. It turns out that the asymptotics sought can be determined by


using the so-called Newtons rule. Consider, for example,
g(u, ) = u3 + u2 + u3 + 2 u2 + 2 u4 + 3 u + 4 u + 5 u5 .
We assign each term of this polynomial to a point of the two-dimensional
lattice with integer co-ordinates, whose abscissa m is equal to the exponent
of  and ordinate n to the exponent of u (see Figure 2.6). Newtons rule
states that, in order to determine the main asymptotics of the stationary
points, it is enough to keep only those terms in (2.58), whose points of (m, n)lattice are the vertices of the minimal convex polygon containing all points
of the polynomials. Further, in order to determine the main asymptotics of

48

CHAPTER 2. PRELIMINARIES

Figure 2.6: (m, n)-lattice and the convex polygon.


the stationary points approaching zero as  0, one has to keep only those
terms whose vertices lie on that portion of the boundary going down from A
to the right and ending in the first lowest point (the point C in Figure 2.6).
One can derive Newtons rule from the variational-asymptotic method.
Indeed, consider terms of g for which m 0, n 3. Since we are interested
in the asymptotics u 0 as  0, all terms except that corresponding to A
can be neglected as small compared with u3 . By the same reasoning all terms
for which m 1, n 2, except that corresponding to B, can be neglected
as small compared with u2 . Finally, all terms for which m 3, n 1,
except that corresponding to C, can be neglected as small compared with
3 u. The remaining terms correspond to the vertices A, B, C, which should
be selected in accordance with Newtons rule. This rule can easily be proved
for polynomials of the general form.
Bending of beams. In the classical Bernoulli-Euler theory of beams, the
energy functional of a bent beam is given by
Z
1 L
Eh4 (u00 )2 dx,
(2.59)
IB [u] =
2 0
where u is the transverse displacement of its central line, E is a positive
constant that depends on the elastic moduli and the geometry of the cross
section, and h and L are the thickness and the length of the beam, respectively. The prime denotes differentiation with respect to x. We assume that
at the ends of the beam the following kinematical boundary conditions are
specified:
u = u0 = 0 at x = 0,
u = uL , u0 = L at x = L.

(2.60)

2.5. VARIATIONAL-ASYMPTOTIC METHOD

49

Among all functions satisfying the constraints (2.60) the true transverse displacement of the central line of the beam is the minimizer of the functional
(2.59).
Timoshenko proposed a refined theory of beams, in which the bending of
a beam is described by two functions, namely the transverse displacement
of its central line u(x) and the rotation of its cross section (x). The latter
corresponds to the angle between the cross sections in the deformed and
initial states. The energy of the Timoshenko beam is given by
Z
1 L
[Eh4 ( 0 )2 + Gh2 ( + u0 )2 ] dx,
(2.61)
I[u, ] =
2 0
where G is a positive constant that depends on the elastic moduli and the
geometry of the cross section. We also assume the following kinematical
boundary conditions at the ends of the beam:
u = 0, = 0 at x = 0,
u = uL , = L at x = L.

(2.62)

Among all functions satisfying the constraints (2.62) the true functions u and
minimize the functional (2.61).
Now we want to show that the Bernoulli-Euler theory of beams can be
regarded as the first approximation (as h 0) of the Timoshenko theory of
beams.
We shall assume that the boundary data uL and L as well as the constants E and G do not change their values as h 0. We also assume that the
minimizer does not considerably change its values on the distance h so that,
in particular, h2 02  2 . We keep the principal terms in the functional
(2.61). Since Eh4 02  Gh2 2 , the principal terms are
Z
1 L
I0 [u, ] =
Gh2 ( + u0 )2 dx.
2 0
We minimize the functional I0 [u, ] taking into account the constraints (2.62).
The minimum of this functional is equal to zero and is attained at functions
u, , which are related to each other by
= u0 .

(2.63)

Thus, the set M0 of the variational-asymptotic scheme consists of pairs of


functions (u, ), where u(x) is an arbitrary function satisfying the boundary
condition (2.60), and (x) is determined from u(x) according to (2.63). The
latter can be interpreted geometrically as follows: in the first approximation

50

CHAPTER 2. PRELIMINARIES

the cross sections remain perpendicular to the deformed central line of the
beam.
If the conditions of the variational-asymptotic scheme are fulfilled (that
means u? is uniquely determined from u0 and the difference I[u0 , ] I[u0 +
u? (u0 , ), ] is small in the asymptotic sense), I[u, ] defined on M0 turns
out to be the first approximation of the functional (2.61). Because on M0
we have I[u, ] = IB [u], the variational problem reduces to the calculation
of u(x) according to the Bernoulli-Euler theory of beams.
Let us check these conditions. For convenience of calculation, let us introduce a new unknown function defined by = + u0 . It has the meaning
of the angle between the deformed cross section and the plane orthogonal to
the deformed central line of the beam. In terms of u and the Timoshenko
energy functional can be rewritten in the form
Z
1 L
[Eh4 (u00 )2 2Eh4 u00 0 + Eh4 02 + Gh2 2 ] dx.
(2.64)
I=
2 0
Fixing u(x) (u(x) is an element of M0 ), we shall determine from u. We
keep in the functional (2.64) the principal terms involving ; they are Gh2 2
and 2Eh4 u00 0 . The term Eh4 02 can be neglected because h2 02  2 .
Thus, we obtain the functional
Z
1 L
(2Eh4 u00 0 + Gh2 2 ) dx.
I1 [u, ] =
2 0
In order to transform this functional we integrate the first term by parts and
use the boundary conditions
= u0

at x = 0,

= L + u0

at x = L.

After this transformation the functional reads


Z
1 L
I1 [u, ] =
[2Eh4 u000 + Gh2 2 ] dx
2 0
2Eh4 u00 (L + u0 )|x=L + 2Eh4 u00 u0 |x=0 .
The terms not standing under the integral sign are known, because u(x)
is given. Therefore the minimization of the last functional reduces to the
minimization of the expression in the square brackets yielding
=

Eh2 000
u .
G

Thus, the angle is uniquely determined from u. Moreover, is of the order


h2 , and its contribution to the energy is small compared with that of the

2.5. VARIATIONAL-ASYMPTOTIC METHOD

51

Bernoulli-Euler energy. Thus, all the criteria of the variational-asymptotic


procedure are fulfilled, and the Bernoulli-Euler theory is indeed the first
approximation to the Timoshenko theory of beams.
The example considered above illustrates the clear physical meaning of
the variational-asymptotic procedure. At the first step one has to minimize
the main part of the energy. When 6= u0 , the energy would be of the
order h2 . This would not be energetically profitable, because otherwise the
energy, as we have shown, would be of the order h4 , which is much smaller
than that of the order h2 . One can easily find the exact solution of the
problem according to the Timoshenko theory of beams and realize that the
results obtained by the variational-asymptotic procedure correspond to the
asymptotic expansion of the solution into the power series of h2 .
Whithams method. In the subsequent chapters the variational-asymptotic
method will mainly be applied to quadratic functionals leading to linear
boundary-value problems. The following example just shows how it works in
the nonlinear case.12
Consider a linear homogeneous partial differential equation
P (t , x )u = 0,

(2.65)

where P (r, s) is a polynom of r and s. Because of linearity this equation


always yields solutions of the type
u = aei(kxt) ,

(2.66)

where a, k, are constants (which may be complex). The necessary and


sufficient condition for (2.66) to be the solution of (2.65) is that k and
satisfy the following equation:
P (i, ik) = 0.

(2.67)

Solutions of the type (2.66) are called harmonic waves, with a being the amplitude, k the wave number, the frequency, and = kx t the phase.
Equation (2.67) is called dispersion relation. Harmonic waves play a fundamental role in the theory of linear differential equations of the type (2.65),
particularly because an arbitrary solution of (2.65) can be presented as superposition of the harmonic waves. The dispersion relation contains all information about equation (2.65); the latter can be uniquely restored from
(2.67).
12

The applications of the variational-asymptotic method in the non-linear shell theory,


in the homogenization of materials with microstructures and in many other problems can
be found in [6].

52

CHAPTER 2. PRELIMINARIES

The question arises: what are the analogues of harmonic waves and the
dispersion relation for non-linear differential equations? It turns out that a
generalization of the harmonic waves in the non-linear case is
u = (a, ),

(2.68)

where a, are functions of x and t, is a periodic function (with the period


2) with respect to , and the characteristic scales of change of the functions
a, ,x , ,t are considerably larger than that of . The function a is analogous
to the amplitude, while ,x and ,t play the role of the wave number and the
frequency, respectively.
We shall now derive the equations for , a, . Let us consider the Euler
equation of the functional
ZZ
(u, u,x , u,t ) dx dt.
(2.69)
I[u] =

We seek a solution of this equation in the form13


u = (, x, t),

(2.70)

where is a function of x and t, is a periodic function with respect to


(with the period 2). We assume that the characteristic scales L and T of
changes of the functions ,x , ,t and (, x, t)|=const are considerably larger
than the characteristic scales l and of changes of the phase . The latter
are defined as the best constants in the inequalities
|,x |

2
,
l

|,t |

2
,

(2.71)

while the former are the best constants in the inequalities


|,xx |

2
,
lL

|,xt |

|x |

,
L

2
,
lT

|,xt |

|t |

,
T

2
,
L

|,tt |

|, | ,

2
,
T
(2.72)

where x = /x with = const, and t = /t with = const.


Therefore it makes sense to call fast variable as opposed to the slow
variables x and t. Thus, in this variational problem we have two small
parameters l/L and /T .
13

The parameter a appears later.

2.5. VARIATIONAL-ASYMPTOTIC METHOD

53

We now calculate the derivatives u,x and u,t . According to (2.70)


u,x = x + | ,x ,

u,t = t + | ,t ,

where the vertical bar followed by denotes the derivative with respect to
the fast variable . Because of (2.71) and (2.72) they can be approximately
replaced by
u,x = | ,x , u,t = | ,t .
Keeping in the functional (2.69) the asymptotically principal terms gives
ZZ
(, | ,x , | ,t ) dx dt.
I0 [] =

Let us decompose the domain into the strips bounded by the lines = 2k,
k = 0, 1, 2, . . .. The integral over can be replaced by the sum of the
integrals over the strips
ZZ
X ZZ
dx dt =
(, | ,x , | ,t ) d d,
(2.73)

where is the coordinate along the line = const, and is the Jacobian of
transformation from x, t to , . In the first approximation one can regard
, ,x and ,t in each strip as independent from . Therefore one obtains the
same problem in each strip at the first step of the variational-asymptotic
procedure: find the stationary points of the functional
Z 2

(, | ,x , | ,t ) d
(2.74)
I0 [] =
0

among periodic functions (). The quantities ,x and ,t are regarded as


constants in the functional (2.74). The Euler equation of this functional is
a second-order ordinary differential equation. Its solutions contain two arbitrary constants. One of them is determined from the condition of periodicity,
the other one can be chosen by fixing the amplitude a as follows: max = a.
Thus, the set M0 of the variational-asymptotic scheme consists of pairs of
the functions a(x, t) and (x, t).
the value of the functional (2.74) at its stationary
Let us denote by 2

point. The quantity is a function of a, ,x and ,t . The sum (2.73), as


l/L 0 and /T 0, can again be replaced by the integral
ZZ
,x , ,t ) dx dt.
(a,
(2.75)

54

CHAPTER 2. PRELIMINARIES

The Euler equations of the functional (2.75) read

= 0,
a



+
= 0.
x ,x t ,t

(2.76)

It turns out that the equation (2.76)1 is the generalization of the dispersion
relation in the non-linear case. This equation approaches the linear dispersion relation when a 0. The peculiarity of the non-linear version of the
dispersion relation is that it depends on the amplitude of waves.
Whithams method [61] can also be applied for variational problems with
more unknown functions of several variables.
Problems
1. Use the variational-asymptotic method to analyze the behaviour of the
stationary points of the function
f (u, v, ) = u2 2u + 4(u 1)v + 2 v 2 + 22 v
as  0.
2. Determine the characteristic scale of change of the following functions:
i) sin(x/l) on the interval [0, 2], ii) exp(x/l) on the interval [0, ).
3. Derive the equilibrium equations for the Timoshenko beam. By integrating these equations show that the transverse displacement u approaches the solution of the equilibrium equations according to the
Bernoulli-Euler theory of beams as h 0.
4. Use Whithams method to derive the non-linear dispersion relation for
the Euler equation of the functional
ZZ
1
1
I[u] =
[ u 2 c(u0 )2 + f (u)] dx dt.
2
2

Part I
Low-frequency vibrations

55

Chapter 3
Elastic shells
3.1

Two-dimensional equations

Geometry of a shell. A shell may be constructed geometrically by means


of its middle surface and its thickness as follows. Let S be a two-dimensional
smooth surface in the three-dimensional Euclidean space E, bounded by a
smooth closed curve S (in the case of closed surfaces S = ). The surface
is described by a vector equation
z = r(x1 , x2 ),
where r is a smooth vector-value function of two variables x1 , x2 . At each
point of the middle surface we restore the segment of length h in the direction
perpendicular to the surface so that its centre lies on the surface. If the length
h is sufficiently small, the segments do not intersect each other and fill some
domain B E (see Figure 3.1). A linear elastic body occupying the domain

Figure 3.1: A portion of a shell.


B in its stress-free undeformed state is called an elastic shell, the surface S
its middle surface, and h its thickness. A plate is the special case of the
shell, whose middle surface is plane. The shell is said to be thin if h is much
57

58

CHAPTER 3. ELASTIC SHELLS

smaller than the characteristic sizes as well as the radius of curvature of the
middle surface.
Let t = r, and n be the tangents and the unit normal to the middle
surface, respectively. It is well known that the surface is determined uniquely
up to a rigid body motion in the Euclidean space if its first and second
fundamental forms

a = t t ,
= t n, = t, n,

satisfying the compatibility conditions of Gauss and Codazzi, are specified


(the fundamental theorem of the theory of surfaces).
Kinematics of a shell. In this chapter we shall study low-frequency vibrations of elastic shells. Regarding a shell as a two-dimensional continuum,
we assume that the shell kinematics is completely specified if its deformed
middle surface z = r(x , t) is known. We define the displacement field u of
the middle surface as follows:
r(x , t) = r(x ) + u(x , t).
We assume that the displacement field u is continuous and as many times
differentiable as required. The first fundamental form of the deformed middle
surface is given by the equation:
a
= (r, + u, )(r, + u, ).
Assuming that u, is small, and neglecting the term u, u, , we get
a
= a + 2t( u,) .
Since the first fundamental form of the surface enables one to determine
distances between points and angles between co-ordinate lines, we define
A = 1/2(
a a ) as measures of extension of the middle surface. Then
the last equation leads to
A = t( u,) = (t( u),) u t, .
Using Gauss formula,

t, =
t + b n,

we obtain
A = u(;) b u,

(3.1)

3.1. TWO-DIMENSIONAL EQUATIONS

59

where u = t u and u = n u are the components of the displacements


referred to the basis {t , n}. The semicolon preceding Greek indices denotes
co-ordinate expressions of the covariant derivatives.
Let us now calculate the change of curvature of the middle surface. The
second fundamental form of the deformed middle surface is given by
b = n
r, .
We define B = b b as measures of bending of the middle surface.
Using the smallness of u to neglect all non-linear terms we have
B = nu, + nr, ,
where n is the change of the normal unit vector. Taking into account the

t = 0 and the smallness of u we can easily show
identities n
n = 1 and n
that
nn = 0, n t = t n = u, n.
Together with Gauss formula this leads to
B = nu; .

(3.2)

There exist alternative measures of bending widely used in the theory of


shells. These are all of the following class:
= B + C A ,
B

(3.3)

where C
are of the order 1/R of smallness. Let us analyze the measures of
bending introduced by Koiter [23] and Sanders [52]. We can rewrite (3.2) in
the form
B = (nu,( );) n,( u,) .

According to Weingartens formula,


n, = b t ,
we have
B = (;) + b( u,) t
= u; + (u b( );) + b( u;) c u.

(3.4)

Here
= nu, = u, + b u ,

(3.5)

60

CHAPTER 3. ELASTIC SHELLS

and b = a b . The raising and lowering of indices of two-dimensional tensors will be done with the help of the surface metrics a and a . We decompose the expression u, t in (3.4) into the symmetric and skew-symmetric
parts
u, t = A + $ ,
where
1
$ = (u, u, ).
2

(3.6)

With (3.4)-(3.6) we obtain


B = + b( A) ,

(3.7)

= (;) + b( $) .

(3.8)

where

One can also choose as the measures of bending of the middle surface,
since the pair A , B can be expressed in terms of A , and vice versa.
If A , B (alternatively, A , ) are known, one can determine the
deformed middle surface uniquely up to a rigid-body motion. Besides, these
measures may be varied independently in the general case. Therefore they
can be referred to as state variables in the theory of linear elastic shells.
Note that while the measures A are dimensionless, B and have the
dimension [L]1 .
Since u depends on t we introduce the following quantities
v = u (velocity),

(acceleration),
a=u

to measure its time rates.


Variational principle. Since the shell is regarded as a two-dimensional
continuum, it is natural to introduce its action functional as follows:
Z t1 Z
( + F u) da dt,
(3.9)
J[u] =
t0

where is the kinetic energy density, the strain energy density, F the
external force, and da the area element. The action functional J[u] is defined
on the space of all admissible displacement fields, where u is assumed to
be continuously twice differentiable and u continuously differentiable. The
kinetic energy density is a quadratic form of u
1
u,

= u
2

(3.10)

3.1. TWO-DIMENSIONAL EQUATIONS

61

where is the mass density per unit area. The strain energy density is a
positive definite quadratic form of the measures of extension and bending of
the middle surface, which, in the general case, is given by
, B ),
= (A
or, alternatively,
, ).
= (A
depends on B (or ), the functional feels the change of the
Since
derivative of u at the boundary S. If the edge of the shell is free, then it is
natural to assume that no constraints are imposed on u at the boundary. If
the edge of the shell is clamped, we assume that J[u] is defined on the space
of admissible displacement fields u satisfying the boundary condition
u = 0,

u, = 0 at Sk ,

(3.11)

where denotes the surface vector normal to the curve S. The last condition of (3.11) expresses the fact that the rotation angle of the edge of the
shell vanishes (clamped edge). Finally, if the edge is fixed, then only the
displacements should vanish
u = 0 at Sk .

(3.12)

We also require that u is given at t = t0 and t = t1


u|t=t0 = u0 ,

u|t=t1 = u1 .

Hamiltons variational principle for elastic shells states that the true displace is the stationary point of the action functional (3.9)
ment field u
J = 0.
In order to derive the equations of motion for the shell let us calculate
the variation of the functional (3.9)
Z t1 Z

u
A
B + Fu) da dt.
(3.13)
J =
(
u
A
B
t0
S
We introduce the following symmetric tensors:

,
A

.
=
B

N =
M

(3.14)

62

CHAPTER 3. ELASTIC SHELLS

From equation (3.13) one can see that N works on the extension of the
middle surface and M on its bending (change of the curvature). Therefore
it is natural to call N (symmetric) membrane stresses, and M bending
moments.
The variations of A and B are equal to
A = r,( u,) ,
B = nu; .

(3.15)

We now substitute (3.14) and (3.15) into (3.13). Assuming the regularity
of N and M and integrating (3.13) by parts with the help of Gauss
theorem, we obtain for the variations vanishing at the boundary S
t1

J =
t0

+ (N r, ); (M n); + F]u da dt = 0.
[
u

Since u is arbitrary inside the region S (t0 , t1 ), we conclude that


= (N r, ); (M n); + F.
u

(3.16)

This is the two-dimensional equation of motion of the shell. Substituting


(3.14) into (3.16), we obtain three differential equations with respect to the
three unknown components of the displacement field u.
Provided the equation (3.16) is fulfilled we reduce the equation J = 0
for the variations not vanishing at the boundary to
Z

t1

t0

{[N r, + (M n); ]u M nu; } da dt = 0.

(3.17)

Now we need to select independent variations at the boundary. The gradient


u; may be resolved in the normal and tangential directions to the boundary
as follows:
u; = u; + u; .
This is due to the identity = + , with being the surface vector
tangential to the curve S (Figure 3.2). Since u; = u/s, we can
integrate by parts that term in (3.17) containing it to get
Z

t1

t0

Z
Ss

{[N r, + (M n); ]u +

(M n)u
s

M nu; } da dt = 0.

3.1. TWO-DIMENSIONAL EQUATIONS

63

Figure 3.2: The vectors , at the boundary.


For the free edge of the shell the variations u and nu; are arbitrary at
Ss ; hence
[N r, (M n); ]

(M n) = 0,
s
M = 0.

(3.18)

These are the free-edge boundary conditions. For the clamped edge, the
kinematical boundary conditions (3.11) should be fulfilled. If the edge is
fixed, (3.12) and (3.18)2 are the boundary conditions at Sk . We assume the
following initial conditions at t = t0 :
u|t=t0 = u0 ,

t=t0 = v0 .
u|

We can also project the equation of motion (3.16) and the boundary conditions (3.18) onto the tangent and normal directions. Applying Weingartens
formula to (3.16) we get

= (T t M;
u
n); + F,

where T is the unsymmetric tensor given by


T = N + b M .
Multiplication of the equation of motion with t and n, respectively, yields
u = T; + b M; + F ,

(3.19)

u = T b M;
+ F,

(3.20)

and
where
F = F t ,

F = F n.

64

CHAPTER 3. ELASTIC SHELLS

Analogously, from (3.18) we obtain the free-edge boundary conditions at Ss


T + b M = 0,

M;
+ (M ) = 0,
s
M = 0.

(3.21)

Note that the number of boundary conditions is equal to four.


is given in terms of the measures A , , one obtains its variation
If
in the form
= n A + m ,
(3.22)
where

,
A

=
.

n =

(3.23)

Using (3.7) one transforms (3.22) to


(

= (n b m) )A + m B .
Thus, the derivation of the equations of motion and the boundary conditions
remains exactly the same, if we set
(

N = n b m) ,
M = m .
In terms of n and m the equations of motion read

u = t
; + b m; + F ,

u = n b m
; + F,

(3.24)
(3.25)

where the unsymmetric tensor t is equal to


[

t = n + b m] .
We use square brackets between indices to denote the operation of alternation: T[] = 1/2(T T ). In the same manner, the following free-edge
boundary conditions at Ss are obtained
t + b m = 0,

m
(m ) = 0,
; +
s
m = 0.

(3.26)

3.1. TWO-DIMENSIONAL EQUATIONS

65

The strain energy. We consider a shell made of an elastic material which


is homogeneous, isotropic and obeys Hookes law. It turns out that, within
the first-order approximation, the 2-D strain energy density is given by
= h[(A )2 + A A ] +

h3
[( )2 + ],
12

(3.27)

where , are Lame constants and


=

.
+ 2

On the basis of equations (3.23) and (3.27), relations between the membrane
stresses and bending moments from one side, and the measures of extension
and bending, from the other side, may be established
n = 2h(A a + A ),
h3
m = ( a + ).
6

(3.28)

These equations are the constitutive equations for the two-dimensional theory
of isotropic elastic shells.
Let F (A) denote the following scalar function of the symmetric tensor
argument:
F (A) = h[(A )2 + A A ].
Then the strain energy density (3.27) can be written in the compact form
= F (A) +

1
F (h).
12

(3.29)

Within the framework of the first-order approximation one can choose


different measures of bending as given by (3.3). The corresponding energies of
bending differ from each other by small terms of the order h/R compared with
unity. Indeed, choose for example the tensor B of (3.7) as the argument
in the energy of bending
= F (A) +

1
F (hB).
12

(3.30)

According to (3.7) the two formulae (3.29) and (3.30) differ from each other
by cross terms of the type h3 bAB. These terms are of the order h/R
compared with unity, since, due to the Cauchy-Schwarz inequality
h3 bAB h2 b(A2 + h2 B 2 ).

66

CHAPTER 3. ELASTIC SHELLS

Therefore, (3.29) and (3.30) are asymptotically equivalent within the firstorder approximation. The proof remains exactly the same for other measures
of bending from (3.3)
Problems
1. Given a spherical shell, whose undeformed middle surface is described
by
z 1 = R sin cos , z 2 = R sin sin , z 3 = R cos .
Let the radial displacement of the shell be u = constant. Find the
measures of extension and bending of the shell.

2. Find the compatibility conditions for A and B (A and ).

3. Determine the displacement field of the middle surface of a shell from


the given fields A and B (A and ) satisfying the compatibility
conditions.
4. Show that the equations of motion and the boundary conditions for
a plate with b = 0 break up into those of longitudinal and flexural
vibrations. Find the corresponding uncoupled equations and boundary
conditions in terms of u and u.
5. Derive the balance equation of energy for a shell. Using this equation,
prove the uniqueness of solutions of the boundary-value problems.

3.2

Asymptotic analysis

Geometry of a shell as three-dimensional body. The shell geometry can conveniently be described in terms of the following curvilinear coordinates x1 , x2 , x3 x (sometimes we drop the index 3) in a domain it
occupies in the stress-free undeformed state
z i (xa ) = ri (x ) + xni (x ),

(3.31)

where z i = ri (x ) are the equations of the middle surface, and ni (x ) are the
cartesian components of the normal vector n to this surface. The co-ordinates
x take values in a domain S R2 , while x [h/2, h/2].
By taking the partial derivatives of (3.31) with respect to x it is easy to
see that the basis vectors e associated with the co-ordinates x are given
by
e = z, = r, + xn, = t ,
(3.32)

3.2. ASYMPTOTIC ANALYSIS

67

where the quantities


= xb
are called shifter. The partial derivative of (3.31) with respect to x yields
e3 = z,x = n.
Therefore the covariant components of the metric tensor are found to be
g = t t = a 2xb + x2 c ,
g3 = 0,

g33 = 1,

(3.33)

where c are the components of the third quadratic form of the middle
surface. The latter can be expressed through a and b according to
c = Ka + 2Hb ,
with H and K the mean and Gaussian curvature, respectively. From (3.33)
one can calculate the determinant of the metric tensor
g = det gab = a(det )2 ,
where a = det a . In the principal coordinates of b we have
det = (1 xb11 )(1 xb22 ) = 1 2Hx + Kx2 .
Therefore
r
=

g
= 1 2Hx + Kx2 .
a

(3.34)

The volume element referred to the co-ordinate system {xa } is equal to


dv = da dx.
Let us find out the contravariant components of the metric tensor g ab as the
inverse matrix of gab . Let be the inverse of . We seek in the form
= C + Db ,
where C, D are the unknown functions. Using the definition of we have
= (C + Db )( xb ) = .
Expanding this we obtain two equations for C and D
C + xKD = 1,

xC + (1 2Hx)D = 0,

68

CHAPTER 3. ELASTIC SHELLS

giving C = (1 2Hx)/ and D = x/. Thus,


=

1
[(1 2Hx) + xb ].

(3.35)

According to (3.33) and (3.35) the contravariant components of the metric


tensor g ab are given by
1
[(1 2Hx)2 a + 2x(1 2Hx)b + x2 c ],
2
= 0, g 33 = 1.

g =
g 3

(3.36)

Let us introduce the characteristic radius of curvature R of the shell middle


surface as the best constant in the following inequalities
q
q
1
1

b b ,
b; b; 2 .
R
R
We assume that
h
 1.
R

(3.37)

Three-dimensional functional. For simplicity of the asymptotic analysis


we shall first analyze the free vibrations of the shell (no external body force
and surface traction). Then the three-dimensional action functional takes
the form (cf. (2.28))
Z t1 Z Z h/2
I=
(T W ) dx da dt.
(3.38)
t0

h/2

In this formula T is the kinetic energy density


1
1
T = w i w i = (a w w + w 2 ),
2
2
with w , w the projections of the displacement vector onto the tangential and
normal directions to the middle surface
w = ti wi ,

w = ni wi .

The quadratic form W corresponds to the strain energy density. If the shell
is made of a homogeneous, isotropic and elastic material, then
1
W = [(g ab ab )2 + 2g ac g bd ab cd ]
2
1
= [(g  + 33 )2 + 2g g   + 4g 3 3 + 2233 ].
2

3.2. ASYMPTOTIC ANALYSIS

69

The components of the three-dimensional strain tensor ab can be calculated


in accordance with (2.15)
i
 = z,(
wi,) = ti( wi,) xb( ti wi,) ,
i
i
23 = z,
wi,x + z,x
wi, = ti wi,x xb ti wi,x + ni wi, ,
i
33 = z,x
wi,x = ni wi,x ,

where the comma preceding x denotes the partial derivative with respect to
the co-ordinate x. In terms of w , w we can rewrite these equations in the
form
 = w(;) b w xb( w;) + xc w,
23 = w,x + w, + b w xb w,x ,
33 = w,x .

(3.39)

In order to express explicitly the dependence of I on the small parameter


h, we introduce the dimensionless co-ordinate
=

x
,
h

[1/2, 1/2].

In terms of x , the components of the strain tensor read


 = w(;) b w hb( w;) + hc w,
1
23 = w| + w, + b w b w| ,
h
1
33 = w| .
h

(3.40)

The vertical bar followed by denotes the partial derivative with respect
to (and not with respect to x ). Thus, the parameter h enters the action
functional (3.38) through the components of the strain tensor ab .
Among terms of W (ab ) the derivatives w| /h and w| /h in 3 and 33
are the main ones in the asymptotic sense. Therefore it is convenient for the
purpose of asymptotic analysis to single out the components 3 and 33 in
the strain energy. To this end let us decompose the strain energy as follows:
W = W + Wk ,
where
Wk = min W.
3 ,33

70

CHAPTER 3. ELASTIC SHELLS

The parts W and Wk are called transverse and longitudinal strain energy, respectively. The latter does not contain 3 and 33 and coincides with
W when the components of the stresses 3 and 33 vanish. Only the transverse strain energy W depends on 3 and 33 . From the definitions of W
and Wk one easily computes
Wk = [(g  )2 + g g   ],
1
W = ( + 2)(33 + g  )2 + 2g 3 3 ,
2
where = /( + 2).
The first step of the variational-asymptotic procedure. Let us assume
that
h
 1,
(3.41)
cs
where is the characteristic scale of change of the function wi in time (see
Section 2.5) and cs the velocity of shear waves. This means that we consider
here only low-frequency vibrations of the shell.
We now keep the formally principal terms in the functional (3.38). Due to
(3.41) the kinetic energy density can be neglected. In the strain energy density only those quadratic terms containing 3 and 33 should be maintained,
where
1
w| ,
h
1

w| .
2h

33
3

One can put g a and 1 as well. Thus, at the first step of the
variational-asymptotic procedure we obtain the following functional
h
I0 =
2

t1

Z Z

1/2

[( + 2)
t0

1/2

1
1
(w| )2 + 2 a w| w| ] d da dt.
2
h
h

It is clear that the functional I0 is negative definite; its maximum is equal to


zero and attained at displacement fields independent of :
w = u (x , t),

w = u(x , t),

(3.42)

where u (x , t) and u(x , t) are arbitrary functions of x and t. Thus, the set
M0 of the variational-asymptotic procedure consists of displacement fields
of the form (3.42).

3.2. ASYMPTOTIC ANALYSIS

71

Next, keeping u (x , t) and u(x , t) fixed, we seek the stationary points


of the functional (3.38) in the form
w = u (x , t) + v (x , , t),
w = u(x , t) + v(x , , t),

(3.43)

in accordance with the variational-asymptotic scheme. By redefining u , u if


needed we can put the following constraints on the functions v , v:
hv i = 0,

hvi = 0.

(3.44)

In this chapter the angle brackets h.i denote the integration over within the
interval [1/2, 1/2]. According to (3.44), u , u describe the mean displacements of the shell.
We substitute (3.43) again into the action functional (3.38) and neglect
all small terms containing v , v. Due to the assumption (3.41) the time rates
of v , v can be removed in the kinetic energy. In the strain energy we can
approximate the strain tensor by
1
33 = v| ,
h
1
23 v| + u, b u ,
h
 A .
Replacing g a and 1 we obtain the following functional:
Z Z
1
h t1
I1 =
h( + 2)( v| + a A )2
2 t0 S
h
1
1
+a ( v| + )( v| + )ida dt,
h
h
where is given by the formula (3.5). It is easy to see that the functional
I1 is negative definite; its maximum is equal to zero and attained at the
following fields:
v = h , v = hA .
Thus, the terms v , v of the asymptotic expansion (3.43) are determined
uniquely through the functions u , u, and the set N of the variationalasymptotic scheme coincides with the set M0 consisting of the displacement
fields u (x , t), u(x , t).
The second step of the variational-asymptotic procedure. At this
step we seek the stationary points of the functional (3.38) in the form
w = u (x , t) h (x , t) + hy (x , , t),
w = u(x , t) hA (x , t) + hy(x , , t),

(3.45)

72

CHAPTER 3. ELASTIC SHELLS

where u , u are regarded as given functions, with and A expressed


through u , u by (3.5) and (3.1), while y and y are unknown functions
which should be determined by the variational-asymptotic procedure. Without limiting generality we force the functions y , y to satisfy the constraints
hy i = 0,

hyi = 0.

(3.46)

According to (3.46) u , u describe the mean displacements of the shell.


The equations (3.45) can also be interpreted as a change of unknown
functions. Indeed, it is obvious that there is a one-to-one correspondence
between the functions w , w and all the collections of functions u , u, y ,
and y satisfying the constraints (3.46). Let A and B be the measures
of extension and bending which are expressed through u , u according to
(3.1) and (3.7). We introduce in every point x of the middle surface the
characteristic scale of change of the deformation pattern l in the longitudinal
directions as the best constant in the following inequalities:
|A, |
max |y, |

where
A = max
S

1
max |y| |,
l
q

A ,

A
,
l

h |B, |

max |y, |

B
,
l

1
max |y| |,
l

(3.47)

q
B = h max B B .
S

This characteristic lengthscale is a function of x and will be called characteristic wavelength (in the longitudinal directions) for short. We suppose
that
h
1
l
in all points of the shell far from its edge.
Let us substitute (3.45) into the action functional (3.38). Due to (3.41)
the time rates of y , y can be neglected in the kinetic energy. In order to
estimate terms in the strain energy let us now calculate the components of
the strain tensor according to (3.40). It is easy to see that
33 =

1
w| = A + y| .
h

We now analyze the components 3 of the strain tensor


23 = + y| + u, hA, + hy,
+ b u hb + hb y + hb hb y| .

(3.48)

3.2. ASYMPTOTIC ANALYSIS

73

Since h is assumed to be much smaller than the characteristic radius of


curvature R of the middle surface as well as the characteristic wavelength
l in the longitudinal directions, one can see that the underlined terms give
small contributions to the strain energy compared with the other terms.
Neglecting them and recalling (3.5), we arrive at the following formula:
23 = y| .

(3.49)

We now turn to  :
 = u(;) h(;) + hy(;) b u
+ hb A hb y hb( u;) + h2 2 b( ;)
h2 b( y;) + hc u h2 2 c A + h2 c y.
By the same reasoning the underlined terms in this formula give small contributions to the strain energy. Neglecting them and remembering the definitions of A and B we get
 = A hB .

(3.50)

According to the equations (3.48), (3.49) and (3.50) the partial derivatives
of y , y with respect to x do not enter the action functional. The functions
y , y do not enter the longitudinal energy. Putting g a and 1 we
obtain
Z Z
h t1
h( + 2)(y| hB )2 + a y| y| i da dt.
I2 =
2 t0 S
Considering this functional under the constraints (3.46), we see that its maximum is equal to zero and is attained at the functions
1
1
y = hB ( 2 ),
2
12

y = 0.

One can continue the iteration procedure, but for the first-order approximation this is not needed.
Average Lagrangian. According to the variational-asymptotic method we
represent the displacement field in the form
w = u (x , t) h ,
1
w = u(x , t) hA + h2 B ( 2 1/12),
2

(3.51)

74

CHAPTER 3. ELASTIC SHELLS

where u (x , t) and u(x , t) are regarded as the unknown functions, while


the quantities , A and B are expressed through u , u by (3.5), (3.1),
and (3.7). We substitute this displacement field into the action functional
(3.38) and integrate over the thickness. If we keep only the principal terms
containing the unknown functions in the average Lagrangian (first-order approximation), then it is enough to put in (3.38) g = a and = 1. On the
displacement field (3.51) the transverse energy vanishes, while the longitudinal energy should be calculated with
 = A hB ,
or, alternatively,
 = A h .
It is easy to see that we end up with the following expression for the average
strain energy:
= hh(A hB )2 + (A hB )(A hB )i
1
= F (A) + F (hB),
12
or the equivalent expression (3.27), if we use instead of B .
Let us now calculate the average kinetic energy. Differentiating the displacement field (3.51) with respect to time and substituting into the threedimensional kinetic energy, we obtain
1
1
1
= hh[u h A + h2 B ( 2 )]2
2
2
12

+ a (u h )(u h )i.
Here is approximated by 1. The cross-terms between u,
u and A , B ,
vanish due to the constraints (3.44) and (3.46). The terms involving A , B
and are small compared with those in the strain energy, due to the assumption (3.41). Neglecting these small terms, we end up with the following
expression:
1
= h(u 2 + a u u ).
2
This is exactly the formula (3.10), with = h.
We turn to the case when tractions ti on the face surfaces S (at =
1/2) do not vanish. Since the deformation of the shell is of the order
 = max(ab ab )1/2 of smallness, the tractions cannot be arbitrary. The order

3.2. ASYMPTOTIC ANALYSIS

75

of smallness of the latter can be estimated from the condition that the work
done by ti should be of the same order as the strain energy, which is 2 h|S|.
The displacements wi should not depend on h when h tends to zero. From
the dimensional analysis it follows that wi (l + R). Therefore ti should
be of the order
h
h
(3.52)
ti = O(( + )).
l
R
We calculate the work done by the tractions ti acting on the face surfaces
Z
Z
Z
i
i
t wi da +
t wi da = [(t w + tw)|=1/2
S+

+(t w + tw)|=1/2 ] da,

(3.53)

where t , t are the components of the tractions referred to the basis {t , n}.
Within the first approximation we can replace in (3.53) by 1. Substituting
(3.51) into (3.53) and using (3.52) to neglect all small terms, we obtain the
following expression for the work done by the surface tractions
Z
({t }u + {t}u) da,
S

where
{t } = t |=1/2 + t |=1/2 ,

{t} = t|=1/2 + t|=1/2 .

Thus, F and F in (3.19) and (3.20) are equal to {t } and {t}, respectively.
Problems
1. Prove the following asymptotic formula
g = a + 2hb + 3h2 2 c + o(h2 /R2 )a .
2. By varying the action functional (3.38), show that the exact 3-D equations of motion for a shell referred to the co-ordinates x , x read
w = ; + ( ),x b ,
w = ; + b + ,x ,
where = , = 3 , = 33 .
3. Find out the order of smallness of the underlined terms in the formulae
for 3 and  .
4. Provide the similar asymptotic analysis for elastic plates. Show that
Kirchhoffs hypotheses are fulfilled for bent plates.

76

CHAPTER 3. ELASTIC SHELLS

3.3

Dispersion of waves in plates

Flexural waves. Consider an infinite plate of thickness h, with z 1 , z 2 the


two-dimensional cartesian co-ordinates in its middle surface. Since b = 0,
the equations of motion (3.24) and (3.25) break up into the equations of
flexural and longitudinal waves.1 Assuming F = 0, we reduce the equation
of flexural waves (3.25) to
h
u = m
(3.54)
, .
For the cartesian co-ordinates {z } the covariant derivatives coincide with
the partial ones, so the comma before indices can be used instead of the
semicolon. The measures of bending become
= (,) = u, .
The constitutive equations (3.28) now read
m =

h3
( + ).
6

Substituting the formulae for m and into the equation (3.54) we obtain
the equation of flexural waves in the form
h
u + D2 2 u = 0,

(3.55)

where 2 u = u,11 + u,22 is the Laplace operator applied to u, and


D = ( + 1)

Eh3
h3
=
6
12(1 2 )

is the flexural rigidity of the plate.


Let us derive the dispersion relation for the harmonic plane waves. We
introduce the dimensionless variables and according to
r
tcs
t
z
=
=
, = .
(3.56)
h
h
h
We can rewrite the equation (3.55) as follows:
u| +
1

+1 2 2
u = 0.
6

(3.57)

Because the plate is infinite, we speak of wave propagation rather than of its vibrations.

3.3. DISPERSION OF WAVES IN PLATES

77

Here and in the sequel the vertical bar preceding indices denotes the derivatives with respect to the corresponding dimensionless variables and, for simplicity, we use the same 2 to denote the dimensionless Laplace operator.
Consider a harmonic plane flexural wave (F-wave) of the form
u = aei(

1 )

(3.58)

where and are the dimensionless wave number and frequency, respectively, the dimensionless wavelength being related to by = 2/. The
appearance of u at successive instants of time would be as shown in Figure
3.3. The propagation velocity of the wave, called phase velocity, is equal to

Figure 3.3: A propagating wave u = a cos( 1 ).


cp = /. Substituting (3.58) in the differential equation (3.57), we obtain
for the various terms
2 u = a2 ei(
2 2 u = a4 ei(

1 )

1 )

u| = a2 ei(

1 )

The resulting dispersion relation is given by


2 =

+1 4
.
6

(3.59)

The graphs of versus for various Poissons ratios are plotted in Figure
3.4. Since the dispersion curves are symmetric about the and axis, only
their portions in the first quadrant of the (, )-plane are shown.
The dimensionless phase velocity of the F-waves is given by
r

+1
cp = =
.
(3.60)

78

CHAPTER 3. ELASTIC SHELLS

Figure 3.4: Dispersion curves of F-waves.


Thus, the phase velocity cp is a linear function of the wave number , and
the flexural waves in plates are dispersive. Equation (3.60) predicts also
unbounded wave velocities for very short wavelengths and high frequencies.
This physically unacceptable behaviour at short waves demonstrates clearly
that the two-dimensional theory, which is derived only for long waves, cannot
in general be applied in the short-wave range.
The other important characteristics of the dispersive waves is the group
velocity [61]. Imagine that u is the sum of two harmonic waves of equal
amplitude a, but slightly different wave numbers and frequencies
u = a cos( 1 ) + a cos[( + ) 1 ( + ) ].
This equation can be transformed into
u = 2a cos

1
1 + ( + ) 1 ( + )
cos
,
2
2

and after neglecting and in the first factor as small compared with
and , we obtain
u = 2a cos( 1 ) cos

1
.
2

Figure 3.5 is a sketch of u for some fixed time, say 1 , which represents a wave
train of wavelength 2/ whose amplitude varies periodically and slowly with
1 over a long wavelength of g = 4/. The amplitude of the wave train
moves with the velocity
cg =

(3.61)

3.3. DISPERSION OF WAVES IN PLATES

79

Figure 3.5: A group of waves.


which is called group velocity. For infinitesimally small (3.61) gives
cg =

d
.
d

Applying this formula to the flexural waves in plates we have


r
d
+1
cg =
=2
= 2cp .
d
6
Longitudinal waves. We now turn to the equation (3.24). With F = 0
we reduce it to
h
u = t
, .

(3.62)

The measures of extension A are expressed through u as follows


A = u(,) .
The constitutive equations between t and A are given by
t = n = 2h(A + A ).
Substitution of the formulae for t and A into (3.62) yields
h
u = h[(2 + 1)u, + 2 u ].

(3.63)

In terms of the dimensionless variables and introduced in (3.56), the


equations of the longitudinal waves (3.63) can be rewritten in the form
u| = (2 + 1)u| + 2 u .

(3.64)

80

CHAPTER 3. ELASTIC SHELLS

Let us seek the solutions of (3.64) in form of the harmonic plane waves
propagating in the direction 1
u = a ei(

1 )

Substituting this into (3.64) and doing the same calculations as in the previous case, we can show that there are two possible types of waves corresponding to
u1 = a1 ei(
u2 = a2 e

1 )

i( 1 )

u2 = 0 L-wave,

u1 = 0 SS-wave.

For the L-waves the dispersion relation reads


2 = 2( + 1)2 ,

(3.65)

while for the SS-waves (symmetric shear waves) we have


2 = 2 .

(3.66)

Thus, in both cases the phase velocities do not depend on the wave numbers.
Therefore, the longitudinal waves in plates are non-dispersive. The dispersion
curves of (3.65) for different Poissons ratios are presented in Figure 3.6

Figure 3.6: Dispersion curves of L-waves.


Comparison with 3-D elasticity. Consider a plate of thickness h, as
shown in Figure 3.7. The cartesian co-ordinate system is selected with z 1 , z 2 plane coinciding with the middle surface of the plate. The face surfaces of the
plate are given by z 3 = h/2 (sometimes we drop the index 3). Assuming

3.3. DISPERSION OF WAVES IN PLATES

81

Figure 3.7: A plate of thickness h.


that the plate is made of a homogeneous isotropic elastic material and the
body force vanishes, we write down the three-dimensional equations of its
motion in terms of the displacements wi in the form (cf. (2.21))
wi = ( + )wj,ji + wi,jj .
The traction-free boundary conditions on the face surfaces z 3 = h/2 read
i3 |z3 =h/2 = [wj,j i3 + (wi,3 + w3,i )]|z3 =h/2 = 0.
Since the co-ordinate system is cartesian, we are allowed to lower all indices,
keeping in mind the summation convention. We non-dimensionalize these
equations by introducing the following variables
r
zi
t
, i = .
(3.67)
=
h
h
The equations of motion and the boundary conditions then take the dimensionless form
wi| = (1 + )wj|ji + wi|jj ,
[wj|j i3 + (wi|3 + w3|i )]|=1/2 = 0,

(3.68)
(3.69)

where = /.
We look for solutions of the boundary-value problem (3.68) and (3.69) in
the form
wi = fi ()ei(

1 )

(3.70)

Substituting (3.70) into the equations of motion (3.68) and boundary conditions (3.69), one can obtain two uncoupled systems.
For waves of the type
w2 = f2 ()ei(

1 )

w1 = w3 = 0 S-waves,

82

CHAPTER 3. ELASTIC SHELLS

we have
f200 + p22 f2 = 0,
f20 |=1/2 = 0,

(3.71)

where the prime denotes the derivative with respect to and


p22 = 2 2 .
The eigenvalue problem (3.71) yields the following eigenfunctions:
f2 = a cos 2n, p2 = 2n, SS(n)-waves,
f2 = a sin (2n + 1), p2 = (2n + 1), AS(n)-waves.
For the SS(0)-waves we have the following dispersion relation:
2 = 2 ,
which is identical with the relation (3.66).
Let us analyze the second case, when
w1 = f1 ()ei(

1 )

w3 = f3 ()ei(

1 )

w2 = 0.

Substitution of this into equation (3.68) gives


f100 + (1 + )if30 + (2 e2 2 )f1 = 0,
e2 f300 + (1 + )if10 + (2 2 )f3 = 0,
where
e

+ 2
=+2=
,

r
e=

=
+ 2

(3.72)

1 2
.
2 2

The boundary conditions (3.69) become


e2 f30 + if1 = 0,
f10 + if3 = 0.

(3.73)

The eigenvalue problem (3.72) and (3.73) admits the symmetric and antisymmetric solutions of the type
f1 () even, f3 () odd (L-waves),
f1 () odd, f3 () even (F-waves).
The characteristic equation of the system (3.72)
2

s + 2 e2 2
(1 + )is

det
=0
(1 + )is
e2 s2 + 2 2

3.3. DISPERSION OF WAVES IN PLATES

83

has four roots given by


s1,2 = ip1 ,
s3,4 = ip2 ,

e2 2 2 ,

p2 = 2 2 .

p1 =

Therefore the symmetric solutions (L-waves) read


f1 = i(a cos p1 + bp2 cos p2 ),
f2 = ap1 sin p1 + b sin p2 ,

(3.74)

where a and b are still unknown constants. The four boundary conditions on
= 1/2 reduce to two equations in a and b
p1
p2
(2 p22 )a cos + 2p2 b cos
= 0,
2
2
(3.75)
p1
p2
2p1 a sin + (2 p22 )b sin
= 0.
2
2
Equating the determinant to zero, we obtain from (3.75) the dispersion relation
4p1 p2 2
tan(p2 /2)
= 2
.
tan(p1 /2)
( p22 )2

(3.76)

This is the Rayleigh-Lamb dispersion relation for the propagation of the


L-waves in plates. From (3.75) we also obtain the amplitude ratio
a
2p2 cos(p2 /2)
(2 p22 ) sin(p2 /2)
= 2
=
.
b
( p22 ) cos(p1 /2)
2p1 sin(p1 /2)
Next, we consider the antisymmetric solutions (F-waves), which are given
by
f1 = i(c sin p1 dp2 sin p2 ),
f2 = cp1 cos p1 + d cos p2 ,

(3.77)

where c, d are unknown constants. The traction-free boundary conditions at


= 1/2 reduce also in this case to two equations for c, d
p1
p2
= 0,
(2 p22 )c sin 2p2 d sin
2
2
(3.78)
p1
p2
2p1 c cos + (2 p22 )d cos
= 0.
2
2
Since the determinant should vanish to guarantee nontrivial solutions, we
derive from (3.78) the following dispersion relation for the F-waves:
tan(p2 /2)
(2 p22 )2
=
.
tan(p1 /2)
4p1 p2 2

(3.79)

84

CHAPTER 3. ELASTIC SHELLS

We also obtain the equation for the ratio c/d


(2 p22 ) cos(p2 /2)
2p2 sin(p2 /2)
c
=

= 2
.
d
( p22 ) sin(p1 /2)
2p1 cos(p1 /2)
Both equations (3.76) and (3.79) can be combined in a single equation
(
1

+1 L-waves,
tan(p2 /2)
4p1 p2 2
=
,
(3.80)
2 2
2
tan(p1 /2)
( p2 )
1 F-waves.

Figure 3.8: Three regions of the , -plane.


Equation (3.80) was first obtained independently by Rayleigh (1888) and
Lamb (1889). However, due to its complexity, the detailed calculations of
branches of the dispersion curves in the , -plane, including also complex
branches, were completed much later. In this section we only study the
asymptotes of the lowest branches in the low-frequency long-wave region
in order to compare this with the dispersion relations (3.59) and (3.65).
Depending on whether (, ) is found in the regions I, II, or III, as shown in
Figure 3.8, we may have p1 , p2 being both imaginary, one imaginary and one
real, or both real, respectively. The dispersion relations will alter their form
accordingly. For low frequencies, the analysis of regions I and II leads to the
result of interest.
In the region I p1 = iq1 , p2 = iq2 and the equation (3.80) becomes

1
4q1 q2 2
tanh(q2 /2)
.
=
tanh(q1 /2)
(2 + q22 )2
In this region there are no roots for L-branches. For F-branches we expand
the hyperbolic tangent
1
tanh x = x(1 x2 + . . .).
3

3.3. DISPERSION OF WAVES IN PLATES

85

Figure 3.9: Dispersion curves of the lowest branch of F-waves: a) 2-D theory:
dashed line and b) 3-D theory: solid line.
Retaining the first two terms, we reduce the dispersion relation to
q2 (1 13 (q2 /2)2 )
(2 + q22 )2
=
.
4q1 q2 2
q1 (1 13 (q1 /2)2 )
Put this in the form
1
1
(2 q22 )2 = 2 q24 q12 (2 + q22 )2 .
3
12
Expanding this and keeping the terms according to Newtons rule, we obtain
the asymptotic formula
2 =

+1 4
+ O(6 ).
6

(3.81)

This result agrees with the relation (3.59) of the two-dimensional plate theory
for long waves. The dispersion curves of the first branch of F-waves according
to the two- and three-dimensional theories for = 0.31 are shown in Figure
3.9. One can see that the difference between them becomes remarkable for
> 0.5. When the exact dispersion curve approaches asymptotically
the straight line = (cr /cs ) from below, where cr is the Rayleigh wave
velocity.
In the region II p1 = iq1 , and we obtain the equation

1
tan(p2 /2)
4q1 p2 2
=
.
tanh(q1 /2)
(2 p22 )2

86

CHAPTER 3. ELASTIC SHELLS

Figure 3.10: Dispersion curves of the lowest branch of L-waves: a) 2-D theory:
dashed line and b) 3-D theory: solid line.
The lowest F-branch has no roots in this region. For the lowest L-branch we
replace tan x x and tanh x x giving
p2
4q1 p2 2
,
= 2
q1
( p22 )2

or (2 p22 )2 = 42 q12 .

(3.82)

Keeping the main terms in (3.82) we find that


2 = 2( + 1)2 + O(4 ).

(3.83)

This is asymptotically equivalent to the the relation (3.65) for long waves. We
show in Figure 3.10 the dispersion curves of the lowest L-branch computed
according to the two- and three-dimensional theories for = 0.31. The good
agreement between these dispersion curves is observed up to the value = 1.
However, the exact curve approaches the line = (cr /cs ) from above as
.
Problems
1. Find the axisymmetric solution to the equation (3.57) of flexural vibrations of an infinite plate subject to the following initial condition:
u(r, 0) = 0,
where (r) is the Dirac -function.

u(r,
0) = (r),

3.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

87

2. Show that the lowest branches of the dispersion curves of F- and Lwaves according to 3-D theory approach the straight line = (cr /cs )
as .
3. Plot the lowest branches of the dispersion curves according to 3-D theory using Mathematica.

3.4

Frequency spectra of circular plates

Flexural vibrations. We investigate the free vibrations of a circular plate


of radius r. For flexural vibrations we look for solutions of the form
u(z , t) = u(z )eit ,
where is a frequency. With (3.55) this leads to
D2 2 u h 2 u = 0.

(3.84)

We could non-dimensionalize this equation using h as a characteristic length.


However the following dimensionless variables
=

z
,
r

4 = 2 hr4 /D =

6 2 r4
( + 1)h2

enable one to see more details in the low frequency spectrum. Now we transform (3.84) to
(2 2 4 )
u = (2 + 2 )(2 2 )
u = 0.

(3.85)

Therefore the solution of (3.85) may be written as


u = u1 + u2 ,
where u1 , u2 satisfy respectively
2 u1 + 2 u1 = 0,
2 u2 2 u2 = 0.

(3.86)

These equations can be solved by the separation of variables. In the polar


co-ordinates %, we have
2 u =

1 2 u
2 u 1 u
+
+
.
%2 % % %2 2

88

CHAPTER 3. ELASTIC SHELLS

Assuming u1 = f1 (%)g1 (), from (3.86)1 we obtain the following differential


equation for g1 ()
g100 = 2 g1 .
Since g1 () is periodic with the period 2, we find that = n, where n is an
integer, and
g1 = sin n, cos n.
Then the equation for f1 (%) reads
n2
1
f100 + f10 + ( 2 2 )f1 = 0.
%
%
This is Bessels equation of order n, which has the following non-singular
solution
f1 = aJn (%).
Combination of g1 and f1 yields
(
sin n
u1 = aJn (%)
cos n

For u2 = f2 (%)g2 () the same results are obtained for g2 (), while for f2 (%)
the modified Bessels equation holds true
f200

n2
1 0
2
+ f2 ( + 2 )f2 = 0.
%
%

The non-singular solution of this equation is given by


f2 = bIn (%).
Combining u1 and u2 we get finally
(
sin n
u = [aJn (%) + bIn (%)]
cos n

(3.87)

Consider the simplest case of the clamped edge, for which the boundary
conditions
u|%=1 =

u|%=1 = 0
%

(3.88)

3.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

89

hold. Substituting (3.87) into (3.88) and equating the determinant to zero,
we obtain the following frequency equation:
Jn ()In0 () In ()Jn0 () = 0.
The three lowest roots nm
table:
m
1
2
3

(3.89)

of (3.89) for n = 0, 1, 2 are given in the following


n=0
3.196
6.306
9.439

n=1
4.611
7.799
10.958

n=2
5.906
9.197
12.402

Figure 3.11: A few normal modes of a clamped circular plate.


The frequencies of vibrations should be calculated by the formula
r
2
nm
h

nm =
,
r2
6(1 )
while the corresponding eigenfunctions are given by
(
sin n
Jn (nm )
unm = [Jn (nm %)
In (nm %)]
In (nm )
cos n

A few of the deformed shapes of the plate are shown in Figure 3.11.
Let us turn to the case of the free edge. The displacement u should then
satisfy the boundary conditions (3.26)2,3 . In the polar co-ordinates %, we
have
1
a11 = 1, a22 = %2 , a11 = 1, a22 = 2 ,
%
1
1
2
= (1, 0), = (0, %), 22
= %, 12
= .
%

90

CHAPTER 3. ELASTIC SHELLS

Calculation of every single term in (3.26)2,3 shows that


h3

( + 1) 2 u,
6
%
2
3
1 u
h 1 u
(

),
m =
6 % % %
h3
2u
m = (2 u + 2 ).
6
%
m
; =

Substituting this into (3.26)2,3 , we obtain the following conditions at the


boundary % = 1:
[( + 1)

3 u
2 u
2
u +

]|%=1 = 0,
%
%2 2
2 u
(2 u + 2 )|%=1 = 0.
%

(3.90)

With u from (3.87) we transform (3.90) to


a{ 3 Jn0 () + (1 )n2 [Jn () Jn0 ()]}
+b{ 3 In0 () + (1 )n2 [In () In0 ()]} = 0,
a{ 2 Jn () + (1 )[n2 Jn () Jn0 ()]}
+b{ 2 In () + (1 )[n2 In () In0 ()]} = 0.

(3.91)

The frequency equation can be expressed by the condition that the determinant of (3.91) vanishes. When n = 0 the frequency equation can be presented
in a simple form
2(1 ) +

I0 ()
J0 ()
0
= 0.
0
J0 ()
I0 ()

(3.92)

The three lowest roots nm of the frequency equation for n = 0, 1, 2 and


= 0.31 have the following values:
m
1
2
3

n=0
0.0
3.004
6.202

n=1
0.0
4.526
7.735

n=2
2.308
5.938
9.185

The first two zero frequencies correspond to the translation u = a and small
rotation u = a sin of the plate without deformation. The frequency according to 21 is the lowest one that is positive. Two deformed shapes of the
plate for n = 0, m = 2 and n = 2, m = 1 are shown in Figure 3.12.

3.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

91

Figure 3.12: Two normal modes of a free circular plate.


Longitudinal vibrations. For a circular plate, referred to an arbitrary
co-ordinate system {x }, the equations of longitudinal vibrations read
h
u = h[(2 + 1)u; + 2 u ].

(3.93)

We look for solutions of these equations in the form


u (x , t) = u (x )eit ,
where is the frequency. The equations (3.93) then reduce to
h 2 u = h[(2 + 1)
u; + 2 u ].

(3.94)

Introducing the dimensionless variables


x
,
=
r

r
= r

we can rewrite (3.94) as follows:


(2 + 1)
u| + 2 u + 2 u = 0.

(3.95)

In accordance with Helmholtzs decomposition theorem we express the vector


field u by
u = | + .
. | ,
where and are two scalar potentials and .
. the mixed components of the
Levi-Civita tensor. Substitution of this decomposition in (3.95) shows that
u will satisfy the latter, if
2 + 2 = 0,
2 + 2 = 0,

(3.96)

where
2 =

2
.
2( + 1)

(3.97)

92

CHAPTER 3. ELASTIC SHELLS

The equations (3.96) can again be solved by the method of separation of


variables. In the polar co-ordinates %, we have
(
(
sin n
cos n
= aJn (%)
, = bJn (%)
.
(3.98)
cos n
sin n
Assume first that the plate is clamped at its edge. It follows then from
the conditions u |%=1 = 0 that
1
+
)|%=1 = 0,
% %
1
(

)|%=1 = 0.
% %
(

(3.99)

Thus, the factors cos n and sin n must be chosen differently for and
in order to satisfy these boundary conditions. Substituting (3.98) into (3.99)
and equating the determinant of the linear equations for a, b to zero, we arrive
at
Jn0 ()Jn0 () n2 Jn ()Jn () = 0.

(3.100)

When n = 0, we have either


= 0,

independent of ,

= 0,

independent of .

or
The frequency equations and their roots are given by
J00 () = 0 1 = 3.832, 2 = 7.016, 3 = 10.174 . . . ,
in the former case, and
J00 () = 0 1 = 3.832, 2 = 7.016, 3 = 10.174 . . . ,
in the latter case, where should be calculated from according to (3.97).
The frequency of vibrations is given by
r

.
=
r
For n 6= 0 we have to solve the equations (3.100) which depend on . The
three lowest roots nm for n = 1, 2, 3 and = 0.31 are given in the following
table:

3.4. FREQUENCY SPECTRA OF CIRCULAR PLATES


m
1
2
3

n=1
3.325
5.374
8.472

n=2
5.172
6.915
9.908

93

n=3
6.719
8.511
11.311

Now consider the free edge. First let us exclude the longitudinal translation and small rotation of the plate which correspond to the zero frequencies.
According to (3.26)1 the following conditions should be posed at the boundary % = 1:
(
u| a + u(|) ) = 0.

(3.101)

Using Helmholtzs decomposition for u in (3.101), we arrive at the boundary


conditions
2
2

)|%=1 = 0,
2
%
%

2
2
2
+ 2 2 2 )|%=1 = 0.
(2
%

%
(2 +

(3.102)

Substituting (3.98) into (3.102) and equating the determinant of the linear
equations in a, b to zero, we find
[Jn0 () + (n2 ( + 1)2 )Jn ()][2Jn0 () + ( 2 2n2 )Jn ()]
+ 2n2 [Jn0 () Jn ()][Jn0 () Jn ()] = 0. (3.103)
For n = 0 the frequency equation (3.103) breaks up into two equations.
When = 0 and is independent of we have
(1 )J00 () + J0 () = 0,

(3.104)

which yields
1 = 2.055,

2 = 5.391,

3 = 8.573 . . .

Here and in what follows = 0.31. When = 0 and is independent of ,


the following equation holds:
2J00 () + J0 () = 0,
whose lowest positive roots are
1 = 5.136,

2 = 8.417,

3 = 11.62, . . .

For n 6= 0 we have to solve the equation (3.103). The numerical values of


the three lowest roots nm of these equations for n = 1, 2, 3 and = 0.31 are
given below

94

CHAPTER 3. ELASTIC SHELLS


m
1
2
3

n=1
0.0
6.004
6.851

n=2
2.346
7.666
8.832

n=3
3.602
9.062
10.82

Problems
1. Derive the frequency equations of flexural vibrations of a circular plate
fixed at its edge.
2. Determine the lowest frequencies in the above-mentioned problem.
3. Plot the lowest root of the equation (3.104) against Poissons ratio .

3.5

Dispersion of waves in cylindrical shells

2-D equations of motion. In this section we investigate the wave propagation in a closed circular cylindrical shell of the thickness h and the radius
R. The shell is assumed to be infinite in extent along its axis. We denote
by x1 the axial and by x2 = R the circumferential coordinate of its middle
surface, respectively (Figure 3.13). The middle surface of the shell is given

Figure 3.13: A closed circular cylindrical shell.


by the equation
z = x1 i1 + R sin

x2
x2
i2 + R cos i3 .
R
R

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

95

Taking the partial derivatives of z, we find the tangent vectors of the middle
surface
t1 = i1 ,

t2 = cos

x2
x2
i2 sin i3 .
R
R

Therefore the unit normal vector to the middle surface is equal to


n=

x2
x2
t1 t2
= sin i2 + cos i3 .
|t1 t2 |
R
R

Calculation of the partial derivatives of n yields


n,1 = 0,

n,2 =

1
t2 .
R

According to the formulae (2.13) and (2.14), the components of the first and
the second quadratic forms read
a = ,
b11 = b12 = 0,

a = ,
1
b22 = .
R

(3.105)

Because of (3.105)1 the operations of raising or lowering of indices do not


affect the components of tensors at all, so we are allowed to lower all indices.

Besides, since the Christoffel symbols


vanish, the covariant derivatives
coincide with the corresponding partial derivatives.
Next, we express the measures of extension A and bending through
the displacements u , u. According to (3.1) and (3.105)
A11 = u1,1 ,

1
A12 = (u1,2 + u2,1 ),
2

A22 = u2,2 +

u
.
R

The quantities from (3.5) and $ from (3.6) are given by


1 = u,1 ,

2 = u,2

u2
;
R

1
$21 = $12 = (u1,2 u2,1 ).
2

Therefore calculating according to (3.8) we obtain


u2,2
,
R

11 = u,11 ,

22 = u,22

12 = u,12

u2,1
1
+
(u1,2 u2,1 ).
2R
4R

96

CHAPTER 3. ELASTIC SHELLS

We now turn to the differential equations of motion of the cylindrical


shell. Assuming in (3.24) and (3.25) F = F = 0 we have
h
u1 = t11,1 + t12,2 ,
1
h
u2 = t21,1 + t22,2 (m21,1 + m22,2 ),
R
n22
m11,11 2m12,12 + m22,22 ,
h
u=
R

(3.106)

where
t11 = n11 ,

t22 = n22 ,
m12
m12
, t21 = n12
.
t12 = n12 +
2R
2R
The constitutive equations for the membrane stresses and bending moments
are given by
u
n11 = 2h[(u1,1 + u2,2 + ) + u1,1 ],
R
u
u
n22 = 2h[(u1,1 + u2,2 + ) + u2,2 + ],
R
R
1
n12 = 2h (u1,2 + u2,1 ),
2
and
h3
u2,2
[(u,11 + u,22
) + u,11 ],
6
R
h3
u2,2
u2,2
m22 =
[(u,11 + u,22
) + u,22
],
6
R
R
h3
u2,1
1
m12 =
[u,12
+
(u1,2 + u2,1 )].
6
R
4R
Substituting these equations into the equations of motion (3.106), we obtain
m11 =

u,1
) + u1,11 ]
R
1 h3
u2,12
1
+ h(u1,22 + u2,12 ) +
[u,122
+
(u1,22 + u2,12 )],
2R 6
R
4R

h
u1 = 2h[(u1,11 + u2,12 +

1 h3
u2,11
1
[u,112
+
(u1,12 + u2,11 )]
2R 6
R
4R
u,2
u,2
1 h3
u2,11
+ 2h[(u1,12 + u2,22 +
) + (u2,22 +
)]
[u,112
R
R
R 6
R
1
u2,22
u2,22
+
(u1,12 + u2,11 ) + (u,112 + u,222
) + u,222
], (3.107)
4R
R
R

h
u2 = h(u1,12 + u2,11 )

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

97

u
u
1
h
u = 2h[(u1,1 + u2,2 + ) + u2,2 + ]
R
R
R
3
h
u2,112

[(u,1111 + u,1122
) + u,1111 ]
6
R
h3
u2,112
1
2
[u,1122
+
(u1,122 + u2,112 )]
6
R
4R
u2,222
u2,222
h3
[(u,1122 + u,2222
) + u,2222
].

6
R
R
We non-dimensionalize these equations by introducing the following variables:
r
t
x
=
, =
.
(3.108)
R
R
Then the equations of motion (3.107) can be written in the matrix form as
Lu = 0,

(3.109)

where u is the displacement vector



u1

u = u2 ,
u
and L is a matrix differential operator. The latter can be presented as the
sum of two operators
L = L D + ? L M ,
where LD is the differential operator according to the Donnell-Mushtari theory, LM is a modifying operator which alters the Donnell-Mushtari theory
to yield the theory of Koiter-Sanders, and ? is the dimensionless thickness
parameter defined by
? =

h2
.
12R2

The Donnell-Mushtari operator takes the form

2( + 1)12 + 22
(2 + 1)1 2
21
2

(2 + 1)1 2
12 + 2( + 1)22
2( + 1)2
LD =
2

21
2( + 1)2
2( + 1)(1 + ? 4 )
+2

98

CHAPTER 3. ELASTIC SHELLS

where 4 = 2 2 , while the modifying operator is given by


1 2

43 1 2
1 22
4 2

3
1 2 9 12 + 2( + 1)22 (3 + 2)12 2
4
4
LM =

2( + 1)23

1 22
(3 + 2)12 2
0
2( + 1)23

Dispersion curves. We seek solutions of the wave equations (3.109) for the
cylindrical shell in the form
(
   
u1
a1 i( 1 ) cos n
=
e
,
u
a3
sin n
(
(3.110)
sin
n
1
u2 = a2 ei( )
,
cos n
where ai are the unknown constants and n takes the values 0, 1, 2, . . .. The
periodic functions of used in (3.110) guarantee that the displacements
are continuous with respect to . Substituting (3.110) into (3.109) and
eliminating the common factor, which is either cos n exp[i( 1 )] or
sin n exp[i( 1 )], we arrive at the following eigenvalue problem
H(, )a = 0,

(3.111)

where H is a 3 3 matrix, whose elements are the (complex) functions of


and . The matrix H is the sum of two matrices HD and ? HM according
to the operators LD and ? LM , respectively. These matrices are given by

(2 + 1)ni
2i
2( + 1)2 n2

+2

2
2
(2 + 1)ni
2( + 1)n
2( + 1)n

HD =
2

2i
2( + 1)n
2( + 1)[1 + ?
(2 + n2 )2 ] 2
and

HM

41 n2

3
ni
4
=

2
n i

34 ni
94 2

2( + 1)n

(3 + 2)n2
+2( + 1)n3

n2 i
2

(3 + 2)n
.
2( + 1)n3

0
2

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

99

The equation (3.111) has non-trivial solutions if and only if its determinant
vanishes
detH = 0.

(3.112)

This is the dispersion relation for the waves (3.110) propagating in the cylindrical shell. One can see that for every fixed n and every real and fixed
(3.112) is a cubic equation with respect to 2 giving six real values of
symmetrically situated about the -axis in the , -plane.

Figure 3.14: Dispersion curves of the AR-waves.


For n = 0 the equation (3.112) breaks up into two equations corresponding to the two uncoupled waves with
u2 independent of ,

u1 = u = 0,

T-waves,

or with
u1 , u independent of ,

u2 = 0,

AR-waves.

For the torsional waves (T-waves) the dispersion relation takes the simplest
form
9
2 = (1 + ? )2 .
4

(3.113)

The T-waves are therefore nondispersive. For the axial-radial waves (ARwaves) we have
[2( + 1)2 + 2 )[2( + 1)(1 + ? 4 ) 2 ] + 4 2 2 = 0.

(3.114)

100

CHAPTER 3. ELASTIC SHELLS

In Figure 3.14 the dispersion curves of (3.114) for = 0.31, h/R = 0.1 are
shown. Note that at = 0 the frequency of radial vibration with u2 6= 0
does not vanish. This frequency, given by
r
p
2
,
c = 2( + 1) =
1
is called the cut-off frequency of the corresponding branch of the dispersion
curves.

Figure 3.15: Dispersion curves for n = 1.


For n 6= 0 the three branches are coupled. However at = 0 we have
again two uncoupled modes of vibrations. The cut-off frequency of the axial
mode of vibration (u1 6= 0) is equal to
p
= n 1 + ? /4,
while those of the radial and circumferential modes of vibrations2 (u2 , u 6= 0)
should be found as the roots of the equation
[2( + 1)n2 (1 + ? ) + 2 ][2( + 1)(1 + ? n4 ) 2 ]
+4( + 1)2 n2 (1 + ? n2 )2 = 0
yielding
2 = ( + 1)[(1 + n2 )(1 + ? n2 )
p
(1 + n2 )2 2? n2 (1 6n2 + n4 )].
2

Or plane strain modes of vibrations.

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

101

In this formula terms containing ?2 were neglected. In Figure 3.15 the dispersion curves of the dimensionless frequencies versus the dimensionless wave
numbers are shown. The parameters chosen for the numerical calculation
are equal to
= 0.31, h/R = 1/10, n = 1.
Comparison with 3-D elasticity.3 Let us regard now the closed circular
cylindrical shell, shown in Figure 3.13, as a three-dimensional homogeneous
isotropic elastic body. Referring it to the cartesian co-ordinates z i and assuming that the body force vanishes, we write down the three-dimensional
equations of motion in terms of the displacements wi (cf. (2.21))
wi = ( + )wj,ji + wi,jj .
The traction-free boundary conditions on the facial surfaces read
wj,j ni + 2w(i,j) nj = 0,

(3.115)

with ni being the components of the outward unit normal vector. We introduce the following dimensionless variables and parameter
t
=
R

zi
= ,
R
i

h
.
2R

The equations of motion then take the dimensionless form


wi| = (1 + )wj|ji + wi|jj ,

(3.116)

where = /, the vertical bar before indices denoting the differentiation


with respect to i .
According to Helmholtzs decomposition theorem the vector field w can
be expressed in terms of a scalar potential and a vector potential as
follows
wi = |i + ijk j|k ,
where ijk are the components of the 3-D Levi-Civita tensor, and where
j|j = ,
3

This material can be omitted in the first reading.

102

CHAPTER 3. ELASTIC SHELLS

with being a function of co-ordinates and time, which can be chosen arbitrarily.4 The equations of motion (3.116) are satisfied if the potentials and
i satisfy the wave equations
| = e2 |ii ,
i| = i|jj .

(3.117)

Introducing the dimensionless cylindrical co-ordinates 1 , and % and denoting by 1 , , % the corresponding physical components of the vector , we
can rewrite the equations (3.117) as
| = e2 ,
1| = 1 ,
1
2 %
,
| = ( 2 ) + 2
%
%
1
2
%| = ( 2 )% 2
,
%
%

(3.118)

where

2
1 2 1
+
+
(% ).
( 1 )2 %2 2 % % %
We look for solutions of the equations (3.118) in the form
=

= f (%) cos n cos( 1 ),


1 = g1 (%) sin n cos( 1 ),
= g (%) cos n sin( 1 ),
% = g% (%) sin n sin( 1 ),

(3.119)

where f , g1 , g and g% are the unknown functions depending only on %, and n


takes the values 0, 1, 2, . . .. The periodic sine and cosine functions of used
in (3.119) guarantee that the displacements are continuous with respect to
. Substituting (3.119) into the equations of motion (3.118) and using the
differential operator notation
Bn,% =

1 d
n2
d2
+

(
1),
d%2 % d%
%2

we obtain
Bn,p1 % [f ] = 0,
Bn,p2 % [g1 ] = 0,
Bn+1,p2 % [g% g ] = 0,
Bn1,p2 % [g% + g ] = 0,
4

This property is called gauge invariance of Helmholtzs decomposition.

(3.120)

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

103

where
p21 = e2 2 2 ,

p22 = 2 2 .

(3.121)

The general solution of (3.120) is given in terms of the Bessel functions J


and Y , or the modified Bessel functions I and K of the argument q1 % = |p1 |%
and q2 % = |p2 |%, depending on whether p1 and p2 , as determined by (3.121),
are real or imaginary. The general solution of (3.120) is
f = aZn (q1 %) + bWn (q1 %),
g1 = a1 Zn (q2 %) + b1 Wn (q2 %),
1
g2 = (g% g ) = a2 Zn+1 (q2 %) + b2 Wn+1 (q2 %),
2
1
g3 = (g% + g ) = a3 Zn1 (q2 %) + 2b3 Wn1 (q2 %),
2

(3.122)

where, for brevity, Z denotes the J or I functions, and W denotes the Y or


K functions. The gauge invariance can now be utilized in order to eliminate
two of the constants entering (3.122). It may be shown that any one of the
three potentials ga , (a = 1, 2, 3) can be set equal to zero. Setting for example
g3 = 0 we obtain
g% = g2 ,

g = g2 ,

and hence the displacement field


w1 = (f g20

n+1
g2 ) cos n sin( 1 ),
%

n
w = ( f + g2 g10 ) sin n cos( 1 ),
%
n
0
w% = (f + g1 + g2 ) cos n cos( 1 ),
%

(3.123)

where the prime denotes the derivative with respect to %.


Substituting the displacement field (3.123) into the traction-free boundary conditions (3.115) and eliminating the common factors, we arrive at the
following conditions at % = 1
n
e2 2 f + 2[f 00 + ( g1 )0 + g20 ] = 0,
%
2n
f
n+1
(f 0 ) (2g100 + p22 g1 ) (
g2 g20 ) = 0,
%
%
%
n
n
n(n
+
1)
2f 0
g1 [ g20 + (
p22 + 2 )g2 ] = 0.
2
%
%
%

(3.124)

104

CHAPTER 3. ELASTIC SHELLS

Together with (3.122) the equations (3.124) at % = 1 lead to a system of


linear equations for a, b, a1 , b1 , a2 and b2 . This system of equations may have
non-trivial solutions if and only if
detCij = 0,

(3.125)

where Cij is a 6 6 matrix, whose first three rows are given by


C11
C12
C13
C14
C15
C16
C21
C22
C23
C24
C25
C26
C31
C32
C33
C34
C35
C36

= [2n(n 1) (p22 2 )2 ]Zn (q1 ) + 21 q1 Zn+1 (q1 ),


= [2n(n 1) (p22 2 )2 ]Wn (q1 ) + 2q1 Wn+1 (q1 ),
= 2n(n 1)Zn (q2 ) 22 nq2 Zn+1 (q2 ),
= 2n(n 1)Wn (q2 ) 2nq2 Wn+1 (q2 ),
= 2q2 2 Zn (q2 ) 2(n + 1)Zn+1 (q2 ),
= 22 q2 2 Wn (q2 ) 2(n + 1)Wn+1 (q2 ),
= 2n(n 1)Zn (q1 ) + 21 nq1 Zn+1 (q1 ),
= 2n(n 1)Wn (q1 ) + 2nq1 Wn+1 (q1 ),
(3.126)
2 2
= [2n(n 1) p2 ]Zn (q2 ) 22 q2 Zn+1 (q2 ),
= [2n(n 1) p22 2 ]Wn (q2 ) 2q2 Wn+1 (q2 ),
= q2 2 Zn (q2 ) 2(n + 1)Zn+1 (q2 ),
= 2 q2 2 Wn (q2 ) 2(n + 1)Wn+1 (q2 ),
= 2nZn (q1 ) + 21 q1 2 Zn+1 (q1 ),
= 2nWn (q1 ) + 2q1 2 Wn+1 (q1 ),
= nZn (q2 ),
= nWn (q2 ),
= nq2 Zn (q2 ) + (p22 2 )2 Zn+1 (q2 ),
= 2 nq2 Wn (q2 ) + (p22 2 )2 Wn+1 (q2 ).

Here
1 = sign(p21 ),

2 = sign(p22 ),

= 1 + .

The remaining three rows of the matrix Cij are obtained from the first three
by substitution of = 1 for .
For waves independent of the angular co-ordinate (n = 0), the determinant in (3.125) breaks into the product of two determinants, so that
D1 D2 = 0,

3.5. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

105

where


C23 C24
,
D1 =
C53 C54


C11

C
D2 = 31
C41
C61

C12
C32
C42
C62

C15
C35
C45
C65


C16
C36
,
C46
C66

(3.127)

and the terms Cij are given by (3.126) with n = 0.

Figure 3.16: Dispersion curves of the lowest branch of AR-waves (n = 0): a)


2-D theory: dashed line, and b) 3-D theory: solid line.
For f = g2 = 0 and
D1 = 0,

(3.128)

one obtains waves involving the displacement w only, i.e., the torsional
waves. It may be ascertained that no roots of (3.128) exist for p22 < 0;
hence equation (3.128) may be reduced to
J2 (q2 )Y2 (q2 ) Y2 (q2 )J2 (q2 ) = 0.

(3.129)

The lowest branch of torsional waves is described by the following dispersion


relation
p22 = 2 2 = 0,
which is asymptotically equivalent to (3.113) of the two-dimensional shell
theory. These T-waves correspond to the following displacement field
w1 = w% = 0,

w = a% sin( 1 ),

106

CHAPTER 3. ELASTIC SHELLS

which describes a rotation of each transverse section of the cylindrical shell


as a whole about its centre. There is no dispersion for waves of this type,
and both the dimensionless phase and group velocities are equal to 1.
The dispersion relation
D2 = 0

(3.130)

corresponds to AR-waves, i.e., waves involving the displacements w1 , w%


which are independent of . A dispersion relation equivalent to (3.130)
has been derived by J. Ghosh (see [2]). The dispersion curves of the first
branch of AR-waves according to the two- and three-dimensional theories
for = 0.31, h/R = 1/10 are shown in Figure 3.16. One can see that the
dispersion curves according to the two- and three-dimensional theories are
practically identical for < 4.

Figure 3.17: Dispersion curves of the three lowest branches of waves for
n = 1: a) 2-D theory: dashed line, and b) 3-D theory: solid line.
For n 6= 0 the waves are coupled, so we have to find the roots of the
equation (3.125). This yields many branches of the dispersion curves in
the , -plane. However, here we study only the three lowest branches in the
low-frequency long-wave region in order to compare them with the dispersion
curves according to (3.112). The dispersion curves of the first three branches
of waves in cylindrical shells according to the two- and three-dimensional
theories for n = 1, = 0.31, h/R = 1/10 are shown in Figure 3.17. Again, in
the long-wave range ( < 4) the difference between them is negligibly small.
Problems

3.6. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

107

1. Find the axisymmetric solution to the equation (3.109) of wave propagation in an infinite closed circular cylindrical shell subject to the
following initial condition at t = 0:
u = u = u = 0,

u = ( 1 ).

2. Plot the phase and group velocities as functions of the wave number
for the AR-waves (n = 0) in a closed circular cylindrical shell with
= 0.31, h/R = 0.1.
3. Plot the dispersion curves of waves for n = 2 in a closed circular cylindrical shell with = 0.31, h/R = 0.1.
4. Compare the lowest branches of the dispersion curves for n = 1, 2, 3.

5. Plot the lowest three branches of the dispersion curves fot n = 2 according to Gazis equation (3.125) using Mathematica.

3.6

Frequency spectra of cylindrical shells

General solutions. We now consider a finite circular cylindrical shell of the


thickness h, the radius R, and the length 2L, referred to the same co-ordinates
as in the previous section. In terms of the dimensionless variables (3.108) the
equations of motion of the cylindrical shell are given in the operator notation
by (3.109). We seek solutions of this equation in the form
(
   
cos n
u1
f
= 1 ei
,
u
f3
sin n
(
sin n
u2 = f2 ei
,
cos n
where f1 , f2 , f3 are functions of 1 and n = 0, 1, 2, . . .. By this we reduce
the partial differential equations (3.109) to a system of ordinary differential
equations, which can symbolically be written as
Mf = 0,
where f is the vector

f1
f = f2 ,
f3

(3.131)

108

CHAPTER 3. ELASTIC SHELLS

and M is a matrix differential operator, which is the sum of two operators


M = MD + ? MM .
The operator MD takes the form

(2 + 1)n1
2( + 1)12 n2
2

(2 + 1)n1
12 2( + 1)n2
MD =

+2

21
2( + 1)n
while the modifying operator is given by
1 2
3n

4n
4 1

3n
9 2
1
2( + 1)n2
4
4 1
MM =

2
n 1
(3 + 2)n12
+2( + 1)n3

21
2( + 1)n
2( + 1)[1 + ? (14
2n2 12 + n4 )] 2

n2 1

(3 +
.
2( + 1)n

0
2)n12
3

The system of equations (3.131) is a linear system of 8th order with constant
coefficients, which depend on n and 2 , where n, being the number of circumferential nodal points, can be chosen arbitrarily. If the boundary conditions
at the two edges of the shell are the same, then it can be shown that general
solutions of (3.131) fall into the two following classes
i) Symmetric solutions:
f1 odd,

f2 , f3 even functions of 1 ,

ii) Antisymmetric solutions


f1 even,

f2 , f3 odd functions of 1 .

We can therefore assume that f1 , f2 , f3 are of the form


(
sin 1
f1 = a
,
cos 1
   (
cos 1
f2
b
=
.
c
f3
sin 1

(3.132)

3.6. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

109

Substituting (3.132) into (3.131) we get a determinantal equation which is


identical to the dispersion relation (3.112) for finding . Note that the determinant is the polynom of 4th degree with respect to 2 , which, for each ,
gives four different roots5 1 , 2 , 3 , 4 up to the sign of . When i satisfy
the dispersion relation, the ratios a : b : c are determined, in terms of i and
, by any two of the equations (3.131). Thus, the general solutions of (3.131)
are of the form
(
4
X
sin i 1
f1 =
ai
,
cos i 1
i=1
(
  X
4  
cos i 1
f2
bi
=
,
f3
ci
sin i 1
i=1
in which the constants ai are arbitrary, but the constants bi , ci are expressed
as multiples of them. The boundary conditions at 1 = l give a system of
four independent homogeneous linear equations in four unknowns ai , and the
condition of vanishing determinant leads to the frequency equation for .
Axisymmetric vibrations. Because of the coupling between the three
functions fi as well as the high order of the dispersion relation, the frequency
spectrum of the cylindrical shell in the general case is rather complicated.
Therefore it makes sense to study the special case n = 0 (axisymmetric
vibrations), for which u1 , u2 , u are independent of and the system (3.131)
breaks up into two sytems. For torsional vibrations with f1 = f3 = 0 we
have
f200 + 2 f2 = 0.
Here the prime is used to denote the derivative with respect to 1 , and the
term 9? /4 is neglected as small compared with 1. This equation is similar
to the equation of vibrations of a string. The frequency spectrum in the case
of free edges is given by
n =

n,
2l

n = 0, 1, 2, . . .

corresponding to the eigenfunctions


f2 = a sin n 1 ,

n odd,

f2 = a cos n 1 ,

n even.

If both edges are fixed, n = 0 should be removed and the eigenfunctions


should be changed.
5

In some ranges of we may have complex conjugate roots.

110

CHAPTER 3. ELASTIC SHELLS

Figure 3.18: Roots of the equation (3.135) for = 0.31, h/R = 0.1: a)
complex roots: combined dashed lines, and b) real or imaginary roots: solid
line.
For axial-radial vibrations with f2 = 0 we have the coupled system of
equations
2( + 1)f100 + 2f30 + 2 f1 = 0,
2f10 + 2( + 1)? f30000 + [2( + 1) 2 ]f3 = 0.

(3.133)

We seek the symmetric solutions of (3.133) in the form


f1 = a sin 1 ,

f3 = c cos 1 .

(3.134)

Substituting (3.134) into (3.133) and equating the determinant to zero, we


see that satisfies the following equation
[2( + 1)2 + 2 ][2( + 1)(1 + ? 4 ) 2 ] + 4 2 2 = 0.

(3.135)

For each there are three different roots of this equation 1 , 2 , 3 up to the
sign of . In (0, ) there are two complex conjugated roots and one real
root, in a small range ( , c ) near the cut-off frequency two imaginary roots
and one real root and in the remaining region of two real roots and one
imaginary root. They are all presented in Figure 3.18.
Thus, the symmetric solutions to (3.133) should have the form
f1 =

3
X
i=1

ai sin 1 1 ,

f3 =

3
X
i=1

ai

i
cos i 1 ,
2i

(3.136)

where ai are unknown constants and i are given by


i = 2( + 1)2i + 2 .

(3.137)

3.6. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

111

Figure 3.19: Frequencies versus l for cylindrical shells with free edges
( = 0.31, h/R = 0.1).
We still have to satisfy the boundary conditions. Let us analyze the three
following variants of them.
i) Free edges. In terms of f1 , f3 the boundary conditions read
( + 1)f10 + f3 = 0,
f300 = f3000 = 0, at 1 = l.

(3.138)

ii) Fixed edges


f1 = f3 = f300 = 0,

at 1 = l.

(3.139)

f1 = f3 = f30 = 0,

at 1 = l.

(3.140)

iii) Clamped edges

Here l = L/R. Substituting (3.136) into the boundary conditions (3.138)(3.140), we obtain the following system of three linear homogeneous equations
3
X

Cij aj = 0,

i = 1, 2, 3.

(3.141)

j=1

The components Cij for the three different types of boundary conditions are
given by:

112

CHAPTER 3. ELASTIC SHELLS

Figure 3.20: Frequencies versus l for cylindrical shells with fixed edges
( = 0.31, h/R = 0.1).
i) Free edges
1
cos j l,
j
C2j = j j cos j l,
C3j = j 2j sin j l,
C1j =

ii) Fixed edges


C1j = sin j l,
j
C2j =
cos j l,
j
C3j = j j cos j l,
iii) Clamped edges
C1j = sin j l,
j
C2j =
cos j l,
j
C3j = j sin j l.
The eigenfrequencies of vibrations should be determined by
detCij = 0.

3.6. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

113

Taking = 0.31, h/R = 0.1, the graphs of the frequencies as functions of


the half-length-to-radius ratio l = L/R (1, 5) for the three above-mentioned
cases are presented in Figures 3.19, 3.20 and 3.21, respectively.

Figure 3.21: Frequencies versus l for cylindrical shells with clamped edges
( = 0.31, h/R = 0.1).
For the calculation of the frequency spectra in the general case n 6= 0
under various boundary conditions and for numerous shell theories, see the
comprehensive survey by Leissa [34].
Shells with shear diaphragms at both edges. Consider the following
type of boundary conditions

11

11

=m

= 0,

u = u2 = 0,
at x = L.

(3.142)

Physically, these conditions can be realized by means of rigidly attaching


a thin, flat, circular plate at each edges. The plates would have considerable stiffnesses in their own planes, thereby restraining the u and u2 components of shell displacement at their mutual boundaries. However, the plates,
by virtue of their thinness, would have very little stiffness in the x1 direction; consequently, they would generate negligible membrane stress n11 and
bending moment m11 . This type of boundary conditions is called shear diaphragm [34].
The closed circular cylindrical shell supported at both ends by shear di-

114

CHAPTER 3. ELASTIC SHELLS

Figure 3.22: Nodal patterns for cylindrical shells supported at both edges
with shear diaphragms.
aphragms admits simple solutions. Indeed, considering
u1 = a1 cos 1 cos nei ,
u2 = a2 sin 1 sin nei ,
u = a3 sin 1 cos nei ,
and choosing
m =

m
,
2l

(3.143)

the boundary conditions (3.142) are readily seen to be satisfied exactly. The
eigenfrequencies can be determined by the equation equivalent to (3.112),
where is replaced by m according to (3.143). This means that the frequencies are determined by the points of intersection of the dispersion curves
with the equidistant vertical lines. Typical nodal patterns of the deformed
shape of the shell are shown in Figure 3.22.
It is worth noting that the same problem can be solved exactly within
the framework of 3-D elasticity. Regarding the closed circular cylindrical
shell with shear diaphragms at both ends as a three-dimensional body, whose
geometry is described in the previous section, we have the following boundary
conditions
w2 = w3 = 0,

11 = 0,

at 1 = l.

(3.144)

3.6. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

115

Figure 3.23: Frequencies of the cylindrical shell with the shear diaphragms
at both edges (n = 1, = 0.31, h/R = 0.1): a) 2-D theory: dashed line, and
b) 3-D theory: solid line.
The last condition can be obtained from the Hamilton variational principle
by letting w1 be varied arbitrarily at the boundaries 1 = l. Considering
the standing waves
w1 = (f g20

n+1
g2 ) cos n cos 1 ei ,

n
w = ( f + g2 g10 ) sin n sin 1 ei ,

n
w = (f 0 + g1 + g2 ) cos n sin 1 ei .

(3.145)

with f, g1 , g2 as described by (3.122) and choosing according to (3.143), one


can check that the boundary conditions (3.144) are satisfied identically. The
frequencies should be determined by Gazis equation (3.125) with replaced
by m ; this means that they are determined by the points of intersection
of the dispersion curves with the equidistant vertical lines. The frequencies
according to the 2-D and 3-D theories are plotted in Figure 3.23 for n =
1, = 0.31, h/R = 0.1. Again, the 2-D shell theory works very well in the
low-frequency long-wave range.
It is interesting to note that certain vibration patterns may have frequencies higher than those related to vibrations which are much more complicated.
For example, the frequency for n = 2 is higher than those for n = 3, 4 for
some m (see [34]).

116

CHAPTER 3. ELASTIC SHELLS


Problems

1. Determine the deformed shape of the cylindrical shell with the free
edges, which vibrates at the lowest frequency as shown in Figure 3.19.
2. Plot the real and complex branches of the dispersion curves for the
cylindrical shell for n = 2.
3. Prove that, for the same n, the fundamental frequency of the shell
clamped at both edges is higher than that of the shell with the same
geometry but with free edges. Compare with that of the shell with
fixed edges.

3.7

Frequency spectra of spherical shells

Two-dimensional equations. A spherical surface distinguishes itself by its


constant curvature. If the two-dimensional co-ordinates are chosen so as the
normal vector to the spherical surface points out of the sphere, the second
fundamental form reads
1
b = a ,
R
where R is the radius of the sphere. This fact will be used to simplify the
equations of motion of a closed spherical shell, whose middle surface has the
radius R.
From (3.1) the measures of extension are obtained
A = u(;) +

1
a u.
R

The measures of bending (3.4) become


B = u;

1
2
u(;) 2 a u.
R
R

We now profit from the freedom in choosing the alternative measures of


bending (3.3). By taking
= B + 2 A ,
B
R
we see that the new measures of bending depend only on u
= u; + 1 a u.
B
R2

(3.146)

3.7. FREQUENCY SPECTRA OF SPHERICAL SHELLS

117

should be expressed in terms of the measures A


The strain energy
; therefore its variation is obtained in the form
and B
A + M
B
.
= N
Using (3.146) we transform this to
+
= (N

2
B .
M )A + M
R

This formula shows that the equations of motion (3.19) and (3.20) remain
exactly the same if we set
+
N = N

2
M ,
R

.
M = M

and M
the equations of motion read
In terms of N
,
h
u = N
;
1
1
.
M
h
u = S 2 M
;
R
R
Together with the constitutive equations
= 2h(A a + A ),
N

3
h
a + B
).
= ( B
M
6
we obtain the following equations in terms of the displacements
u,
1
)u + 2(2 + 1) ],
2
R
R
2(2 + 1)
2u
h2

u = {
(u; + ) + [( + 1)2
R
R
6
2 + 1
2
+
](2 + 2 )u},
R2
R

u = [(2 + 1)u; + (2 +

(3.147)

with 2 u = a u; the Laplace operator applied to u. In deriving (3.147)


we use the following rules for interchanging the order of covariant derivatives
1
u ,
R2
1
= 2 2 u + 2 2 u,
R
u; u; =

(u; );

which are related to the fact, that the intrinsic geometry of the spherical
surface is non-Euclidean.

118

CHAPTER 3. ELASTIC SHELLS

Solutions of the equations of motion. We seek solutions of the equations


(3.147) in the form
u = u eit ,

u = ueit ,

where is a frequency. Introducing the following dimensionless quantities


r
x

h2

=
, = R
, ? =
,
R

12R2
we can rewrite equations (3.147) as follows
(2 + 1)
u| + (2 + 1)
u + 2(2 + 1)
u| + 2 u = 0,
2(2 + 1)(
u| + 2
u) + 2? [( + 1)2

(3.148)

+(2 + 1)](2 + 2)
u 2 u = 0.
Next, let us decompose the vector field u according to
u = | + .
. | ,

(3.149)

where and are two scalar functions. By substituting (3.149) into (3.148)
and observing the rules for interchanging the order of covariant derivatives,
we get
[2( + 1)2 + 2(2 + 1)
u + (2 + 2)]|
2
2
+.
. [ + ( + 2)]| = 0,
2(2 + 1)(2 + 2
u) + 2? [( + 1)2 + (2 + 1)]
(2 + 2)
u 2 u = 0.

(3.150)

The solutions of (3.150) fall into one of the following classes


i) Tangential vibrations with = u = 0. The equation for reads
2 + (2 + 2) = 0.

(3.151)

ii) Radial-tangential vibrations with = 0. The equations for and u


are coupled
2( + 1)2 + 2(2 + 1)
u + (2 + 2) = 0,
2(2 + 1)(2 + 2
u) + 2? [( + 1)2 + (2 + 1)]
(2 + 2)
u 2 u = 0.

(3.152)

3.7. FREQUENCY SPECTRA OF SPHERICAL SHELLS

119

Frequency spectra
Class i) The analysis of the equations of motion has been done so far for
arbitrary surface co-ordinates. Now we use the spherical co-ordinates, in
terms of which the middle surface is given by
z 1 = R sin cos ,

z 2 = R sin sin ,

z 3 = R cos .

The dimensionless components of the metric tensor read


a11 = 1,

a12 = a21 = 0,

a22 = sin2 .

The non-vanishing Christoffel symbols are


2
2
12
= 21
= cot ,

1
22
= sin cos ,

and the 2-D Laplace operator 2 takes the form


2

1 2
= 2 + cot
+
.

sin2 2
2

Thus, equation (3.151) becomes

1 2
2
+
cot

+
+ (2 + 2) = 0.
2
sin2 2

(3.153)

It is well known that (3.153) has regular solutions everywhere on the sphere
if and only if
2 + 2 = j(j + 1),

j = 0, 1, . . . .

In this case the eigenfunctions are given by


Pjm (cos ) cos m,

Pjm (cos ) sin m

(3.154)

where Pjm (x) are Legendres functions [14]. The eigenfunctions (3.154) are
called spherical harmonics, tesseral harmonics for m < j and sectoral harmonics for m = j. These functions are periodic with respect to the angles
and with periods and 2, respectively.
The axisymmetric modes are determined by
= Pj (cos ),

= u = 0,

where Pj (x) are Legendre polynomials. The displacements are calculated


according to (3.149). For each integer j > 0 there is one axisymmetric mode

120

CHAPTER 3. ELASTIC SHELLS

and j non-axisymmetric modes, given by m = 1, . . . , j, corresponding to the


spherical harmonics of order m. The frequency , however, is independent
of m, the wave number in peripheral direction. This seemingly paradoxical
situation can be explained as a consequence of spherical symmetry.
Class ii) The vibrations of this class are determined by the equations (3.152).
For the closed shell and u are spherical harmonics. Supposing
u = aPjm (cos ) cos m,

= bPjm (cos ) cos m,

and substituting these into (3.152), we find the following system of linear,
homogeneous equations for a and b
2(2 + 1)a + [2 + 2 2( + 1)j(j + 1)]b = 0,
{2? [2 j(j + 1)][( + 1)j(j + 1) + 2 + 1]
2 + 4(2 + 1)}a 2(2 + 1)j(j + 1)b = 0.

(3.155)

When j = 0, the second equation yields


2 = 4(2 + 1)(1 + ? ).
This mode corresponds to a purely radial vibration with P0 = 1.
In the general case (j > 0) the frequency is determined from the condition
that the determinant of (3.155) vanishes. For each value of j there are two
roots of this equation. For j = 1, 2, 3, h/R = 0.1 and = 0.31 they are given
by the following table
n
1
2

j=1
0.0
3.375

j=2
1.2
4.647

j=3
1.482
6.2

Also in this case the frequencies are independent of the wave number m.
The vibrations of closed spherical shells were first studied by Lamb, who
solved the 3-D problem by adapting his earlier derived results on the vibrations of elastic spheres to the case of a shell bounded by two concentric
spherical surfaces.
Problems
1. Find the radial vibration of the spherical shell subject to the initial
conditions
u(t = 0) = 0, u(t
= 0) = v0 .

3.7. FREQUENCY SPECTRA OF SPHERICAL SHELLS

121

2. Plot the frequencies determined by (3.155) against Poissons ratio


(0, 0.5).
3. Determine the frequency spectrum for the half of the spherical shell
with the free edges, which is an approximate model of the church bell.

4. Try to solve the 3-D problem of vibrations of the closed spherical shell
and to compare with the solution obtained by the 2-D shell theory.

122

CHAPTER 3. ELASTIC SHELLS

Chapter 4
Elastic rods
4.1

One-dimensional equations

Geometry of a rod. Take an arbitrary plane figure S that is connected


(but need not be simply connected). By moving this figure along a smooth
curve c(x) so that c(x) always remains orthogonal to S and cuts its centroid,
we fill some domain B in the three-dimensional Euclidean point space E
(Figure 4.1). A linear elastic body occupying the domain B in its stress-free

Figure 4.1: A rod.


undeformed state is called an elastic rod, the curve c(x) its central line, and
S its cross section.
Let z = r(x) be the equation of the central line, with x being the arclength. Then the unit tangent vector to the central line can be obtained
as
t = r0 (x).
123

124

CHAPTER 4. ELASTIC RODS

We choose two unit vectors t1 , t2 orthogonal to each other and to t so that


they are rigidly mounted with the cross section1 and move together with S
as it moves along the central line, and so that t1 , t2 and t form a positiveoriented triad. Because t1 , t2 and t are unit vectors and mutually orthogonal,
we have
t t = 1, t t = , t t = 0.
Differentiating the first identity one can see that t0 is orthogonal to t. We
express t0 as a linear combination of t :
t0 = t .

(4.1)

Since the dual basis coincides with t1 , t2 and t, the raising or lowering of
indices does not affect the values of tensor components. Therefore we can
place indices of vectors and tensors referred to the basis {t , t} arbitrarily, at
our convenience and in accordance with the summation rule. In exactly the
same manner we can see that t0 is orthogonal to t , so it can be expressed
as a linear combination of t and e.
. t
t0 = t + $e.
. t ,

(4.2)

where e are the two-dimensional permutation symbols (e11 = e22 = 0, e12 =


e21 = 1). The quantities 1 ,2 and $ are called the curvatures and torsion
of the rod. When $ 6= 0, the rod is called naturally twisted.
Kinematics of a rod. We assume that the kinematics of a rod, regarded as
a one-dimensional continuum, is completely specified if its deformed central
line z = r(x, t) and triad t , t are known. We define the displacements u of
the central line as follows:
r(x, t) = r(x) + u(x, t).
The function u(x, t) is assumed to be continuous and as many times differentiable as required. The arc-length s of the deformed central line satisfies
the following equation
 2
ds
= (r0 + u0 )(r0 + u0 ).
dx
Assuming that u0 is small and neglecting the term u0 u0 , we get
 2
ds
= 1 + 2r0 u0 .
dx
1

For instance we can choose them co-directional to the axes of symmetry of the cross
section, if these exist.

4.1. ONE-DIMENSIONAL EQUATIONS

125

We define = 1/2((ds/dx)2 1) as elongation of the central line. Then the


last equation leads to
= r0 u0 = t u0 = u0 + u ,

(4.3)

where u = t u and u = t u are the components of the displacements


referred to the basis {t , t}.
Concerning the transformation of the triad we have
t = R t ,

t = R t,

with R(x, t) being an orthogonal rotation tensor (RT R = 1). Now let us
assume that the triad rotates on a small angle, which is typical for the linear
theory. In this case one can show that
R 1 + W,
where W is a skew-symmetric tensor (WT = W). Denoting by (x, t) the
axial vector associated with W, we present the transformation rule for the
triad in the form
t = t + t ,

t = t + t.

(4.4)

Referring to the basis {t , t}


= t + t,
one can see that its components are uniquely determined from the displacements. Indeed, (4.4) and the formula t t = e.
. t yield
t = t + e t .
At the same time
t = (r0 + u0 ) dx (t + u0 )(1 ) t(1 ) + u0 .
ds
Comparison between the last two equations leads to
0
e.
. = t u

or
0
. 0
.
= e.
. t u = e. u $u + e. u.

(4.5)

The component is not determined from the displacements. It describes an


additional degree of freedom associated with a rotation of the cross section

126

CHAPTER 4. ELASTIC RODS

about the tangent vector t of the curve c(x). We assume that (x, t) is
continuously differentiable.
For the deformed triad the formulae similar to (4.1) and (4.2) hold true:
t0 dx =
t ,
ds

t0 dx =

t + $e
.
. t .
ds

It is therefore natural to introduce the measures of bending and twist in the


following manner
ds


dx
ds
=
$
$ =
dx
=

= t t0 ,

(4.6)

1 0
e t t $.
2

(4.7)

If , and are known, one can determine the deformed central line and
triad uniquely up to a rigid-body motion. Besides, these measures may be
varied independently in the general case. Therefore they can be referred to
as state variables in the theory of linear elastic rods. It is easy to see that
while is dimensionless, and have the dimension [L]1 .
Keeping in mind (4.1), (4.2) and (4.4) we calculate t0
t0 = t0 + e 0 t + e t0 = t + e 0 t + e t + $ t .
At the same time
t = t e.
. t + e t .
Substituting the last two formulae into (4.6) and neglecting all small terms
we obtain
0
.
= e.
(4.8)
. $ + e. .
Using the formula (4.5), one can rewrite (4.8) also in the following form:
0
.
= (t u0 )0 + e.
. $t u + e. .

Analogously,
0
0
.
0
. 0
.
t0 = t0 e.
. t e. t + e. t + e. t
.
.
. 0
.
. 0
= t + e.
. $t e. t + e. t + e. t + e. ( t + $e. t ),

and therefore
= 0 + = 0 e t u0 .
Since u and depend on t we introduce the following quantities:
v = u (velocity),

(acceleration),
a=u
,

(4.9)

4.1. ONE-DIMENSIONAL EQUATIONS

127

to measure their time rates. The latter quantities describe the angular velocity and acceleration of the torsional motion of the rod.
Variational principle. Consider the following 1-D functional:
Z t1 Z L
( + F u + Q) dx dt,
J[u, ] =
t0

(4.10)

where is the kinetic energy density, the strain energy density, F the
external generalized force, Q the external generalized twisting moment, and
L the total length of the central line. The action functional J[u, ] is defined on the space of admissible functions u and , where u is assumed to
be continuously twice differentiable, while the remaining functions u, are
continuously differentiable. The kinetic energy density is a quadratic form
of u and .
We assume that
1
1
u + 2 ,
= u
(4.11)
2
2
where is the mass density per unit length, and is a constant to be determined later. The strain energy density is a positive definite quadratic form
of , and , which, for the rod made of a homogeneous, isotropic elastic
material, turns out to be given by
1
= (E|S| 2 + EI + C2 ).
2

(4.12)

The coefficient E is Youngs modulus, |S| is the area of the cross section,
EI describe the flexural rigidity and C is the torsional rigidity.
Requiring u and to be specified at t = t0 and t = t1
u|t=t0 = u0 ,
|t=t0 = 0 ,

u|t=t1 = u1 ,
|t=t1 = 1 ,

Hamiltons variational principle for elastic rods states that the true displace and rotation correspond to the stationary points of the action
ments u
functional (4.10)
J = 0.
To obtain the consequences from this variational principle let us calculate
the variation of the functional (4.10)
Z t1 Z L
u +
J =
{
u
T tu0 M [(t u0 )0 + e $t u0
t0

+ e ] M (0 e t u0 ) + Fu + Q} dx dt. (4.13)

128

CHAPTER 4. ELASTIC RODS

Here the following notations are introduced

= E|S|,

M =
= EI ,

= C.
M=

T =

(4.14)

We call T tension, M bending moments, and M twisting moment. These


equations, expressing the relationship between T, M , M and , , , are
regarded as the constitutive equations for elastic rods.
We first assume that both the edges of the rod are clamped. This means
u = 0,

t u0 = 0,

= 0,

at x = 0 and x = L. The functions u and are prescribed at t = t0 , t1 ;


therefore
u = 0,

= 0 at t = t0 , t1 .

Integrating (4.13) by parts, we transform the equation J = 0 to


Z t1 Z L
+ (T t)0 + (M0 t )0 (e M $t )0 + (e M t )0
{[
u
t0

+F]u + (
+ M 0 e M + Q)} dx dt = 0.
Since u and may be given independently and since they are arbitrary
inside the region (0, L) (t0 , t1 ), we conclude that
= (T t)0 + (M0 t )0 (e M $t )0 + (e M t )0 + F,
u
= M 0 e M + Q.

(4.15)

These are the one-dimensional equations of motion of the elastic rod. Substituting the constitutive equations (4.14) into (4.15), we obtain four differential
equations with respect to the four unknown functions u, .
The projections of the equations of motion (4.15) onto the directions of
t and t lead to
u = T 0 + M0 e M $ + F,
0
0
0
u = T + M00 e.
. [M $ + (M $) (M ) ]

(M $ M )$ + F ,
0

= M e M + Q,

(4.16)

4.1. ONE-DIMENSIONAL EQUATIONS

129

where F = F t and F = F t .
Boundary-value problems. Practically there are three kinds of boundary
conditions for the rod:
i) Clamped edge
u = 0,

t u0 = 0,

= 0.

ii) Fixed edge


u = 0,

= 0,

but t u0 may vary arbitrarily.


iii) Force and moments applied to the edge (or free edge). In this case we
should add to I a term associated with the work of a force and moments
and obtain the boundary conditions by varying u and arbitrarily at
the edge of the rod.
In case ii) the additional boundary condition is
M = 0.
If there are no force and moments applied to the edge of the rod (free edge),
the following conditions should be satisfied:
T = 0,

M = 0,

M0 = 0,

M = 0.

(4.17)

Note that the number of the boundary conditions at each edge of the rod is
equal to six. We also assume the following initial conditions at t = t0 :
u|t=t0 = u0 ,
|t=t0 = 0 ,

t=t0 = v0 ,
u|
|
t=t0 = 0 .

130

CHAPTER 4. ELASTIC RODS


Problems

1. Consider a rod whose undeformed central line coincides with the z 3 -axis
z 1 = 0,

z 2 = 0,

z 3 = x,

and whose undeformed unit vectors t1 , t2 rigidly mounted with its cross
section (for example, ellipse with symmetry axes in the directions t1 ,
t2 ) are given by
x
x
t1 = (cos , sin , 0),
a
a

x
x
t2 = ( sin , cos , 0).
a
a

Find the curvatures and torsion of the rod.


2. Determine the deformed central line and the deformed triad of a rod
from given functions , , .
3. Show that the equations of motion for a naturally straight and untwisted rod with = 0, $ = 0 break into the system of uncoupled
equations for longitudinal, flexural and torsional motions. Find the
corresponding equations in terms of u, u and .
4. Derive the balance equation of energy for a rod. Using this equation,
prove the uniqueness of solutions of the boundary-value problems.

4.2

Asymptotic analysis

Geometry of a rod as three-dimensional body. The most convenient


way to describe the rod geometry is to introduce the following curvilinear
co-ordinates x1 , x2 , x3 x (sometimes we drop the index 3) in a domain
occupied by a rod in its stress-free undeformed state
z i (xa ) = ri (x) + ti (x)x ,

(4.18)

where z i = ri (x) is the equation of the central line, and ti (x) are cartesian
components of the vectors t (x). The co-ordinates x take values in a connected domain S R2 . We assume that the point
co-ordinates x = 0
R with
coincides with the centroid of the domain S, so S x da = 0.
By taking the partial derivatives of (4.18) with respect to xa it is easy to
see that the basis vectors ea associated with the co-ordinate system xa are
given by

e = t , e3 = (1 + x )t + $e.
(4.19)
. x t .

4.2. ASYMPTOTIC ANALYSIS

131

Therefore the components of the metric tensor and its determinant are found
to be
g = ,

g3 = $e x ,

g33 = (1 + x )2 + $2 x x ,
g = det gab = (1 + x )2 .

(4.20)

Let us find out the contravariant components of the metric tensor g ab as the
inverse matrix of gab . According to the definition of g ab we have
g g + g 3 g3 = ,
g 3 g + g 33 g3 = 0,

(4.21)

g 3 g3 + g 33 g33 = 1.
From the second equation of (4.21) it follows that
g 3 = g 33 g3 .
Substituting this into the third equation of (4.21) and making use of the
formula for g3 , we get
g

33

1
=
,
(1 + x )2

$e.
. x
=
.
(1 + x )2

(4.22)

Now the remaining components g can be calculated from the first equation
of (4.21):
.
$2 e.
. e. x x

g = +
.
(4.23)
(1 + x )2
Note that the dual basis ea = g ab eb depends on x , x as is evident from (4.22)
and (4.23).
Denote by L the total length of the central line c(x), by h the diameter of
S (the longest distance between two points of S), and by R the best constant
in the following inequalities:
| |

1
,
R

|0 |

1
,
R2

|$|

1
,
R

|$0 |

1
,
R2

which is called the characteristic radius of curvatures and torsion of the rod.
When h/L  1 and h/R  1, the rod is said to be thin.
Three-dimensional functional. For simplicity of the asymptotic analysis
we shall first consider the free vibrations of the rod (no external body force

132

CHAPTER 4. ELASTIC RODS

and surface traction). Then the three-dimensional action functional takes


the form (cf. (2.28))
Z t1 Z L Z

(T W ) g da dx dt.
I=
(4.24)
t0

In the functional (4.24) T is the kinetic energy density


1
1
T = w i w i = (w w + w 2 ),
2
2

(4.25)

where w = ti wi , w = ti wi are the projections of the displacement vector w


onto the directions t , t. The quadratic form W is the strain energy density.
If the rod is made of a homogeneous isotropic elastic material, then
1
1
W = (g ab ab )2 + g ac g bd ab cd = (g  + 2g 3 3 + g 33 33 )2
2
2

33
+ g g   + 2g g 3 3 + 2g 3 g 3 3 3 + 2g 3 g 3  33
+ 2g g 3  3 + 2g 3 g 33 3 33 + (g 33 33 )2 . (4.26)
The components of the three-dimensional strain tensor ab can be calculated
in accordance with (2.15)
i
 = z,(
wi,) = ti( wi,) = w(,) ,
1 i
i
3 = (z,
wi,x + z,x
wi, )
2
1
i
= [ti wi,x + (1 + x )ti wi, + $e.
. t x wi, ]
2
1

.
= [w,x w $e.
. w + (1 + x )w, + $e. x w, ],
2
i
i
33 = z,x
wi,x = (1 + x )ti wi,x + $e.
. t x wi,x

.
= (1 + x )(w,x + w ) + $e.
. x (w,x w $e. w ),

where the comma preceding x denotes the partial derivative with respect
to the co-ordinate x. In this section, components of two-dimensional vectors
and tensors like w , , e are always referred to the orthonormal basis {t }
and its dual basis {t }. This is very covenient for computations, since t ,
in contrast to e , do not depend on x and coincide with t . Consequently,
the raising or lowering of Greek indices with the help of the Kronecker delta
does not affect the components of these two-dimensional tensors.
It is convenient to express explicitly the dependence of I on the small
parameter h. To this end we introduce the dimensionless co-ordinates
=

1
x .
h

4.2. ASYMPTOTIC ANALYSIS

133

The domain of is denoted by S R2 ; it does not depend on h and has


the diameter 1. The integral over S will be denoted by h.i. In terms of , x
the components of the strain tensor read
1
w(|) ,
(4.27)
h
1
1

.
= [w,x w $e.
. w + (1 + h )w| + $e. w| ],
2
h

.
= (1 + h )(w,x + w ) + h$e.
. (w,x w $e. w ).

 =
3
33

We use the vertical bar preceding a Greek index to denote the differentiation
with respect to .
The variational-asymptotic analysis will be considerably simplified if we
neglect from the beginning small terms of the energy in the asymptotic sense.
For the first order approximation, in which terms of the order h/R are neglected as small compared with 1, it is easy to see from (4.20), (4.22) and
(4.23) that we can replace g in (4.24) by 1 and g ab in (4.26) by ab . The
strain energy then takes the form
= 1 (  + 33 )2 +   + 2 3 3 + (33 )2 .
W
2

(4.28)

In this approximate expression for the strain energy, among terms of ab the
derivatives w| /h and w| /h in  and 3 are the principal ones in the
asymptotic sense. Therefore it is convenient to single out the components
 and 3 in the strain energy. To this end let us decompose the strain
energy (4.28) as follows
=W
+ W
+ W
k,
W

(4.29)

where
k = min W
,
W
 ,3

and
= min(W
W
k ).
W


k depends only on 33 and coincides with W

The longitudinal energy W

when the stresses and vanish; the shear energy W depends only
is called the transverse energy. From the
on 3 ; the remaining part W

134

CHAPTER 4. ELASTIC RODS

k, W
and W
one can easily compute
definitions of W
k = 1 E(33 )2 , E = (3 + 2) ,
W
2
+

W = 2 3 3 ,
= ( 1 + )( + 33 )(, , ),
W
2
where = /2( + ) is Poissons ratio, and the short notation (, , )
means the preceding expression with the indices , replaced by , .
The first step of the variational-asymptotic procedure. Let us assume
that
h
 1,
(4.30)
cs
where is the characteristic scale of change of the function wi in time (see
Section 2.5) and cs the velocity of shear waves. This means that we consider
here only low-frequency vibrations of the rod.
We now keep the formally principal terms in the action functional (4.24).
Due to (4.30) the kinetic energy density can be neglected. In the strain
energy density only those quadratic terms containing  and 3 should be
maintained, where
1
w(|) ,
h
1
3
w| .
2h
Thus, the first step of the variational-asymptotic procedure yields the following functional:
Z Z


t1

0 idx dt,
hW

I0 = h2

t0

where

1
2
0 = 1 [ 1 (w|
) + w(|) w(|) + w| w| ].
W
2
h 2
2

Since W0 is the positive definite quadratic form with repect to w(|) and w| ,
it is clear that the functional I0 is negative definite; its maximum is equal to
zero and attained at the following displacement fields:
w = u (x, t),

w = u(x, t),

(4.31)

where u (x, t) and u(x, t) are arbitrary functions of x and t. Thus, the set
M0 of the variational-asymptotic procedure consists of displacement fields
of the form (4.31).

4.2. ASYMPTOTIC ANALYSIS

135

Next, keeping u (x, t) and u(x, t) fixed, we represent the displacement


field in the form
w = u (x, t) + v ( , x, t),
w = u(x, t) + v( , x, t),
in accordance with the variational-asymptotic scheme. Without limiting generality we require the functions v , v to satisfy the constraints
hv i = 0,

hvi = 0.

(4.32)

By this we can interpret u , u as the mean displacements of the rod.


We again neglect all small terms containing v , v in the action functional
(4.24). Due to (4.30) the time rates of v , v can be neglected in the kinetic
energy. In the strain energy we keep only those quadratic terms containing
 and 3 , which are approximated by
1
v(|) ,
h
1 1
1 1
.
( v| + u0 u $e.
. u ) = ( v| + e. ),
2 h
2 h

 =
3

with being given by (4.5). It can be shown that the component 33 is
approximated by 33 , but at this stage we shall neglect the cross-terms
 33 and 3 33 in the strain energy. These cross-terms will be taken into
account in the next step of the variational-asymptotic procedure. Thus, we
get from (4.24) the following functional
Z t1 Z L
2
1 idx dt,
I1 = h
hW
t0

where
1
2
1 = 1 [ 1 (v|
) + v (|) v(|) + (v| + he.
W
. )( )].
2
h 2
2
It is easy to see that the functional I1 is negative definite; its maximum is
equal to zero and attained at the following fields:
v = he (x, t) ,

v = he.
. .

These fields describe the rotation of the cross-section, where the function
(x, t) represents an additional degree of freedom. It will be shown that the
next step brings no more degrees of freedom, and the set N of the variationalasymptotic procedure consists of functions u , u and .

136

CHAPTER 4. ELASTIC RODS

The second step of the variational-asymptotic procedure. At this


step we represent the displacement field of the rod in the form
w = u (x, t) he (x, t) + hy ( , x, t),

w = u(x, t) he.
. (x, t) + hy( , x, t),

(4.33)

where u , u, are regarded as given functions and are expressed through


u , u by (4.5), while y and y are unknown functions which should be determined by the variational-asymptotic procedure. Without limiting generality
we can force the functions y , y to satisfy the constraints
hy i = 0,

hyi = 0.

According to these equations u , u describe the mean displacements of the


rod. The additional degree of freedom associated with the rotation of the
cross section enables one to put also the following constraint on y
hy| ie = 0.
It means that describes the mean rotation of the cross section.
The equations (4.33) can also be interpreted as a change of unknown functions. Indeed, it is obvious that there is a one-to-one correspondence between
the functions w , w and all the collections of functions u , u, , , y , and y
satisfying the above mentioned constraints. Let , , be the measures of
elongation, bending and twist, which are expressed through u , u and by
(4.3), (4.8) and (4.9). We introduce the characteristic scale of change of the
deformation pattern l as the best constant in the following inequalities:





d

, d , d ,


l
dx
l
dx
l
l
(4.34)
q
q
1
1
|
|
|y,x | max y| y
|y,x | max y| y ,
l
l
where the quantities with bars denote their amplitudes. This characteristic
lengthscale is a function of x. We suppose that
h
1
l
in all points of the rod far from its ends.
Let us substitute (4.33) into the action functional (4.24). Due to (4.30)
the time rates of y , y can be neglected in the kinetic energy. In order to

4.2. ASYMPTOTIC ANALYSIS

137

estimate terms in the strain energy let us now calculate the components of
the strain tensor according to (4.27). It is easy to see that
 =

1
w(|) = y(|) .
h

(4.35)

We now analyze the components 3 of the strain tensor


1

3 = [u0 he 0 + hy,x (u he.


. + hy)
2

.
$e.
. (u he + hy ) + (e. + y| )
.
+ (he.
. + hy| ) + $e. (he + hy| )].

Since h is assumed to be much smaller than the characteristic radius of curvatures and torsion R as well as the lengthscale of change of the deformation
pattern l, one can see that the underlined terms give small contributions
to the strain energy compared with the other terms. Neglecting them and
rearranging the remaining terms, we arrive at the following formula:
1
.
3 = [u0 u $e.
. u e. + y|
2

.
he 0 + h e.
. h e. ].
Recalling equations (4.5) and (4.9) we reduce this formula to
1
3 = (y| he ).
2

(4.36)

We now turn to 33


0

33 = (1 + h )(u0 he.


. + hy,x + u
0
0
h e + h y ) + h$e.
. [u he + hy,x

(u he.
. + hy) $e. (u he + hy )].

By the same reasoning the underlined terms in the formula for 33 give small
contributions to the strain energy. We neglect them and rearrange the remaining terms to obtain
0

33 = u0 + u he.
. h e

0
.
+h$e.
. (u u $e. u ).

Remembering the definitions of and we get finally


33 = + h .

(4.37)

138

CHAPTER 4. ELASTIC RODS

According to the equations (4.35), (4.36) and (4.37) the partial derivatives
of y , y with respect to x do not enter the action functional. Thus, the
determination of y , y reduces to the minimization problems for every fixed
x of the following functionals:
1
(4.38)
[y] = h2 h (y| he )( )i,
2
1
[y ] = h2 h ( + 2 )[y(|) + ( + h )]
2
[, , ]i,
(4.39)
under the constraints
hyi = 0,

(4.40)

and
hy i = 0,

hy| ie = 0.

(4.41)

The functionals (4.38) and (4.39) represent the shear and transverse strain
energies, integrated over the cross section of the rod. They are positive
definite and convex, so the existence of their minimizers y , y is guaranteed.
We shall see in the next section that the minimum of (4.39) is equal to zero,
while that of (4.38) is equal to 1/2C2 , with C the torsional rigidity.
Average Lagrangian. Assume that we have solved the cross section problem (4.38) and (4.39) and determined the functions y , y minimizing the
functionals (4.38) and (4.39). In accordance with the variational-asymptotic
scheme we represent the displacement field by (4.33) with y , y replaced by
y , y, but now the functions u , u, should be regarded as unknown functions. We substitute this displacement field into the energy functional (4.24)
and integrate over the cross section. If we keep only the principal terms
containing these unknown functions in the average Lagrangian, then, within
the first-order approximation, it is enough to put g = 1. On the chosen displacement field the transverse strain energy vanishes, the shear energy takes
the minimum value which is equal to 1/2C2 , while the longitudinal energy
should be calculated with
33 = + h .
The average strain energy per unit length of the rod is thus given by
1
= (E|S| 2 + EI + C2 ),
2

4.2. ASYMPTOTIC ANALYSIS

139

where
I = h4 h i
describe the moments of inertia of the cross section.
Let us calculate now the average kinetic energy. Differentiating the displacement field (4.33) with respect to time and substituting into the threedimensional kinetic energy, we obtain
1
+ hy)
2
= h2 h[(u he.
.
2
+ (u he
+ hy )( )]i.
As before g is approximated by 1. The cross-terms between functions u,
u
and ,
, y,
y vanish due to the constraints (4.32), (4.40) and (4.41). The
terms involving , y , y are small compared with those in the strain energy,
due to the assumption (4.30). Neglecting these small terms, we obtain the
following expression:
1
1
= |S|(u 2 + u u ) + h4 h i 2 .
2
2
This is exactly the formula (4.11), with = |S| and
=

h4 h i
.
|S|

Thus, = Ip , where Ip = I is the polar moment of inertia of the cross


section.
We now consider the case when tractions ti on the lateral surface S
(0, L) do not vanish. The work done by the tractions ti acting there, within
the first approximation, is expressed by
Z LZ
h
ti wi ds dx,
(4.42)
S

Substitutwhere ds is the dimensionless element of length of the contour S.


ing (4.33) into (4.42) and neglecting all small terms, we obtain the following
formula for the work done by the surface tractions
Z L
(F u + F u + Q) dx,
0

where
i

F =h

t ds,

Q = h

Problems

e t ds.

140

CHAPTER 4. ELASTIC RODS

1. Find the asymptotic expansions for the co- and contravariant components of the metric tensor in terms of h.
2. Provide the similar asymptotic analysis for an elastic rod, which is
straight and untwisted in its natural state.
3. Find the moments of inertia for the elliptical and rectangular cross
sections.

4.3

Cross section problems

Average transverse strain energy. We want first to show that the minimum of the functional (4.39) is equal to zero. We look for the minimizer in
the form
h i
1

y = a (
),
2
|S|

(4.43)

where a are the components of a third-rank tensor, which is symmetric


with respect to the last two indices. The derivatives of y are given by
y| = a .
It is clear that y from (4.43) satisfy the constraints (4.41). We choose a
so that
1
a() = (a + a ) = h .
2

(4.44)

For arbitrary third-rank tensors which are symmetric with respect to the last
two indices, one can check directly that the following identity
a = a() + a() a()
holds true. Consequently, the solution of (4.44) reads
a = h( + ).

(4.45)

On the field (4.43),(4.45) the functional (4.39) vanishes identically.


Average shear strain energy. We now turn to the minimization problem
(4.38),(4.40). First of all it is easy to see that the constraint (4.40) does
not affect the minimum value of (4.38), because the latter is invariant with
respect to the change of unknown function y y + c, with c a constant.

4.3. CROSS SECTION PROBLEMS

141

By such the change one can always achieve the fulfilment of the constraint
(4.40). Varying the functional (4.38) we obtain the following equation
2 y = 0,
and boundary condition
(y| he ) = 0,
Thus, function y satisfies the
where is the unit outward normal to S.
so-called Neumann problem.
One can also derive a variational problem, which is dual with respect
to the problem (4.38). Using the Legendre transformation we represent the
integrand in (4.38) as follows
1
1
(y| he )( ) = sup[p (y| he )
p p ].
2
2
p
We can now reformulate the problem (4.38) as
inf = h2 inf suphp y| p he
y

1
p p i.
2

(4.46)

Assume that we can interchange the order of taking inf and sup in (4.46).
Then
inf = h2 sup(hp he
y

1
p p i + inf hp y| i).
y
2

It is easy to see that


(
0
if p| = 0, p = 0,
inf hp y| i =
y
otherwise.
Therefore
inf = h2
y

sup hp he
p (4.48)

1
p p i,
2

(4.47)

where the notation p (4.48) means that the supremum in the right-hand
side of (4.47) should be sought among p satisfing the following constraints

p| = 0 in S,

p = 0 at S.

(4.48)

142

CHAPTER 4. ELASTIC RODS

This dual variational problem can be rewritten in a slightly different manner.


Express the solutions of (4.48)1 in terms of a scalar potential
p = e | .

(4.49)

The second constraint in (4.48) then becomes


e | = | =

= 0 at S.
ds

This equation may be integrated along the boundary yielding

= const at S.

(4.50)

In the case of a simply connected cross section we can choose for example
Substituting (4.49) into the functional (4.47) we obtain
= 0 at S.
inf = h2 sup h| h
y

(4.50)

1
| | i.
2

(4.51)

The maximizer of the dual problem satisfies Poissons equation with a constant value for the potential at the boundary.
The minimizer y of as well as the maximizer of (4.51) are proportional to h. Therefore the torsional rigidity C can be calculated by
C = h4 inf h (
y| e )( )i
y

= h4 sup h2|
(4.53)

1
| | i,

= const at S,

(4.52)

(4.53)

where y = y/h, = /h. We can use the dual variational problems (4.52)
to obtain lower and upper bounds for the torsional rigidity.
Elliptical cross section. Let us find out the torsional rigidity for a rod
with an elliptical cross section
c 1,
where c are the components of a positive definite symmetric second-rank
tensor. We seek the minimizer y of (4.52) so that the following equation
(
y| e ) = ae c

(4.54)

4.3. CROSS SECTION PROBLEMS

143

is fulfilled, where a is still an unknown constant. Solving (4.54) with respect


to y| we have
y| =

1 .
ae c + e .
.

(4.55)

The constant a can now be determined from the compatibility condition


1
e (
y| )| = e ( ae.
c + e ) = 0,
.
yielding
a=

2
,
c

c = c .

Integrating (4.55) we obtain


1 2
h i
y = ( e c + e )(
).
2 c
|S|

(4.56)

From (4.55) one can see that y satisfies 2-D Laplaces equation. Taking into
account that

(c ) c

at the boundary of the ellipse, we readily check that the boundary condition
(
y| e ) = 0
is fulfilled. Thus, y given by (4.56) is indeed the minimizer of the functional
(4.52). Substitution of (4.56) into the functional (4.52) gives
C=

4
c c I ,
c2

(4.57)

where I = h4 h i are the moments of inertia of the cross section.


In the co-ordinate system associated with the principal axes of the ellipse
h2
= 2,
b1

c22

1
I11 = b31 b2 ,
4

I22

c11

h2
= 2 , c12 = 0,
b2
1 3
= b2 b1 , I12 = 0,
4

where b1 , b2 are the half-lengths of the major and minor axes. Substituting
these formulae into (4.57) we get finally
C=

b31 b32
4
= 1 ,
2
2
b1 + b2
(I )

144

CHAPTER 4. ELASTIC RODS

where (I 1 ) is a tensor inverse to I .


It is easy to see that (4.56) is also the solution to the problem (4.52) for
a hollow elliptical cross section
2 c 1,

< 1.

Similar calculations give


C=

(1 4 )b31 b32
.
b21 + b22

Rectangular cross section. Let us consider a rectangular cross section


of width a and height 1. We assume for definiteness that a < 1. We try
to determine the function y satisfying Laplaces equation in the rectangular
area and the boundary conditions
y|1 (a/2, 2 ) 2 = 0,

y|2 ( 1 , 1/2) + 1 = 0.

We introduce a new function


f = y 1 2 .
It is straightforward that f satisfies Laplaces equation and the boundary
conditions
f|1 (a/2, 2 ) = 0,

f|2 ( 1 , 1/2) = 2 1 .

(4.58)

Let us assume the solution in form of an infinite series


1

f ( , ) =

cn Xn ( 1 )Yn ( 2 ).

(4.59)

n=0

We require each term of the series (4.59) to satisfy Laplaces equation. By


the separation of variables we can show that
Y 00
Xn00
= n = kn2 ,
Xn
Yn
where kn is a constant to be determined. Thus the factors in the product
solution are
(
(
sin kn 1 ,
sinh kn 2 ,
Xn =
,
Y
=
,
n
cos kn 1 ,
cosh kn 2 .

4.3. CROSS SECTION PROBLEMS

145

Since the derivatives of f are even in both 1 and 2 at the boundaries, we


seek solutions with the same property. Thus we present the solution in the
form
f ( 1 , 2 ) =

cn sin kn 1 sinh kn 2 .

n=0

Because of the boundary conditions (4.58)1 we have


1
cos(kn a) = 0
2
or
kn =

(2n + 1)
,
a

n = 0, 1, . . . .

(4.60)

The boundary conditions (4.58) now require that


f|2 ( 1 , 1/2) =

cn kn sin kn 1 cosh kn

n=0

1
= 2 1 .
2

(4.61)

By virtue of (4.60) the sine functions are orthogonal in the interval a/2
1 a/2; thus, multiplying both sides of (4.61) by sin kn 1 and integrating
over this interval, we obtain
cn =

8a2 (1)n
.
3 (2n + 1)3 cosh(kn /2)

Consequently

8a2 X (1)n sinh kn 2


y = 3
sin kn 1 + 1 2 .
n=0 (2n + 1)3 cosh(kn /2)

(4.62)

One can show that the series (4.62) converges uniformly in 1 and 2 . The
graph of y for a = 1 is plotted in Figure 4.2.
We now use the solution (4.62) to calculate the torsional rigidity. Note
that from (4.52) we can also derive the following formula
C = h4 h (
y| e )e i,
where y is now the solution. Substituting (4.62) into this formula and intergrating over the cross section, we obtain
C = ch4 a3 ,

146

CHAPTER 4. ELASTIC RODS

Figure 4.2: The graph of y.

Figure 4.3: Graph of the function c(a).


where

1 64a X tanh((2n + 1)/2a)


c= 5
.
3
n=0
(2n + 1)5
The graph of c versus a, a [0, 1], is plotted in Figure 4.3.
Problems
1. For the elliptical cross section, find the potential .
2. Establish the following upper bound for the torsional rigidity
4
C 1
(I )
for arbitrary cross sections.
3. Show that among all cross sections with the same polar moment of
inertia Ip = I the circle has the biggest torsional rigidity.

4.4. DISPERSION OF WAVES

4.4

147

Dispersion of waves

Equations of motion of straight rods. In this section we investigate the


wave propagation in an infinite rod, which is straight and untwisted in its
natural undeformed state so that = $ = 0. We choose the triad t , t,
with t co-directional with the principal axes of the moments of inertia I .
Assuming that all the external forces and moments are equal to zero, we
write down the equations of motion of the rod as follows
u = T 0 ,
u = M00 ,
Ip = M 0 .
The measures of elongation, bending and twist of the rod are given by
= u0 ,
= u00 ,
= 0 .
Taking the constitutive equations (4.14) into account, we see that the waves
propagating in the rod fall into three classes: longitudinal, torsional and
flexural. We now study the dispersion of waves of each individual class.
Longitudinal waves. The equation of motion assumes the simple form

u = Eu00 .
Introducing the dimensionless variables
s
t E
,
=
h

(4.63)

x
,
h

we rewrite the equation of the longitudinal waves (4.63) as follows


u| = u00 .

(4.64)

For the harmonic waves of the type


u = aei( )
the dispersion relation reads
2 = 2 .

(4.65)

148

CHAPTER 4. ELASTIC RODS

Thus, the longitudinal waves in rods are similar to the L-waves in plates;
they are both non-dispersive. Of course, this is true only in the low-frequency
long-wave range.
Torsional waves. The equation of motion is identical in form to that of
longitudinal motion and reads
Ip = C00 .

(4.66)

Therefore the previous results can be used with some minor changes. For
example, in terms of the dimensionless variables
s
C
x
t
, = .
=
h Ip
h
the dispersion relation for the harmonic waves
= aei( )
is given by (4.65). Also in this case the waves are non-dispersive.
Flexural waves. Let the rod vibrate in one of the principal planes, so that
only u1 6= 0. The equation of motion reads
|S|
u1 = EI11 u0000
1 ,

(4.67)

where I11 is the corresponding principal moment of inertia. We now introduce


the following dimensionless variables and constant
s
x
I11
t E
, = , 2 =
.
=
h
h
|S|h2
We present equation (4.67) in the dimensionless form
u1| + 2 u0000
1 = 0.
Consider the harmonic flexural wave of the form
u1 = aei( ) .
The resulting dispersion relation is then given by
2 = 2 4 .

(4.68)

4.4. DISPERSION OF WAVES

149

Figure 4.4: A rod of circular cross section.


Comparison with the 3-D theory. Let us regard the infinite rod of circular cross section, shown in Figure 4.4, as a three-dimensional isotropic elastic
body. Referring the rod to the cartesian co-ordinates z i , we can write down
the three-dimensional equations of its motion in terms of the displacements
wi in the form
wi = ( + )wj,ji + wi,jj .
The traction-free boundary conditions on the lateral surface read
wj,j ni + 2w(i,j) nj = 0,

(4.69)

where ni are the components of the outward unit normal vector. We introduce the following dimensionless variables
r
zi
t
, i = .
=
h
h
The equations of motion then take the dimensionless form
wi| = (1 + )wj|ji + wi|jj ,

(4.70)

where = /.
In a similar manner as for cylindrical shells we use Helmholtzs decomposition theorem to express the vector field w in terms of a scalar potential
and a vector potential
wi = |i + ijk j|k ,

150

CHAPTER 4. ELASTIC RODS

It can be shown that the potentials and i satisfy the wave equations
| = e2 |ii ,
i| = i|jj .

(4.71)

Introducing the dimensionless cylindrical co-ordinates 1 , , and % we can


rewrite the equations (4.71) as
| = e2 ,
1| = 1 ,
1
2 %
| = ( 2 ) + 2
,
%
%
1
2
%| = ( 2 )% 2
.
%
%
where
=

(4.72)

2
1 2 1
+
+
(% ).
( 1 )2 %2 2 % % %

We look for solutions of the equations (4.72) in the form


= f (%) cos n cos( 1 ),
1 = g1 (%) sin n cos( 1 ),
= g (%) cos n sin( 1 ),
% = g% (%) sin n sin( 1 ).

(4.73)

Substituting (4.73) into the equations of motion (4.72) and using the differential operator notation
Bn,% =

d2
1 d
n2
+

(
1),
d%2 % d%
%2

we obtain
Bn,p1 % [f ] = 0,
Bn,p2 % [g1 ] = 0,
Bn+1,p2 % [g% g ] = 0,
Bn1,p2 % [g% + g ] = 0,

(4.74)

where
p21 = e2 2 2 ,

p22 = 2 2 .

(4.75)

The general solution of (4.74) is given in terms of the Bessel functions J,


or the modified Bessel functions I of the argument q1 % = |p1 |% and q2 % =

4.4. DISPERSION OF WAVES

151

|p2 |%, depending on whether p1 and p2 , as determined by (4.75), are real or


imaginary. The Bessel functions Y or K must be discarded because of their
singular behaviour at the origin. The general solution of (4.74) is
f = aZn (q1 %),
g1 = a1 Zn (q2 %),
1
g2 = (g% g ) = a2 Zn+1 (q2 %),
2
1
g3 = (g% + g ) = a3 Zn1 (q2 %),
2

(4.76)

where, for brevity, Z denotes the J or I functions. The gauge invariance


can be used to eliminate two of the constants entering (4.76). Setting for
example g3 = 0 we obtain
g% = g2 ,

g = g2 .

and hence the displacement field


w1 = (f g20

n+1
g2 ) cos n sin( 1 ),
%

n
w = ( f + g2 g10 ) sin n cos( 1 ),
%
n
w% = (f 0 + g1 + g2 ) cos n cos( 1 ).
%

(4.77)

Substituting the displacement field (4.77) into the traction-free boundary conditions (4.69) and eliminating the common factors, we arrive at the
following conditions at % = 1/2
n
e2 2 f + 2[f 00 + ( g1 )0 + g20 ] = 0,
%
n+1
2n
f
(f 0 ) (2g100 + p22 g1 ) (
g2 g20 ) = 0,
%
%
%
n
n
n(n + 1)
2f 0
g1 [ g20 + (
p22 + 2 )g2 ] = 0.
%
%
%2

(4.78)

Together with (4.76) the equations (4.78) at % = 1/2 lead to a system of linear
equations for a, a1 and a2 . This system of equations may have non-trivial
solutions if and only if
detCij = 0,

(4.79)

152

CHAPTER 4. ELASTIC RODS

where Cij is a 3 3 matrix, whose elements are given by


C11
C12
C13
C21
C22
C23
C31
C32
C33

= [2n(n 1) (p22 2 )/4]Zn (q1 /2) + 1 q1 Zn+1 (q1 /2),


= 2n(n 1)Zn (q2 /2) 2 nq2 Zn+1 (q2 /2),
= q2 Zn (q2 /2)/4 (n + 1)Zn+1 (q2 /2),
= 2n(n 1)Zn (q1 /2) + 1 nq1 Zn+1 (q1 /2),
= [2n(n 1) p22 /4]Zn (q2 /2) 2 q2 Zn+1 (q2 /2),
= q2 Zn (q2 /2)/4 (n + 1)Zn+1 (q2 /2),
= nZn (q1 /2) + 1 q1 Zn+1 (q1 /2)/2,
= nZn (q2 /2)/2,
= nq2 Zn (q2 /2)/2 + (p22 2 )Zn+1 (q2 /2)/4.

(4.80)

Here
1 = sign(p21 ),

2 = sign(p22 ).

For axially symmetric waves with f, g1 , g2 independent of the angular


co-ordinate (n = 0), the determinant in (4.79) breaks into the following
product


C11 C13


(4.81)
C31 C33 C22 = 0,
where the terms Cij are given by (4.80) with n = 0.
For f = g2 = 0 and
C22 = 0,

(4.82)

one obtains waves involving the displacement w only, i.e., torsional waves.
It may be ascertained that no roots of (4.82) exist for p22 < 0; hence the
latter may be reduced to
q2 J0 (q2 /2) 4J1 (q2 /2) = 0.

(4.83)

The lowest branch of torsional waves is described by the following dispersion


relation
p22 = 2 2 = 0,
which is asymptotic equivalent to (4.65) of the one-dimensional rod theory.
These T-waves correspond to the following displacement field
w1 = w% = 0,

w = a sin( 1 ),

4.4. DISPERSION OF WAVES

153

Figure 4.5: Dispersion curves of the lowest branch of L-waves (n = 0): a)


1-D theory: dashed line, and b) 3-D theory: solid line.
which describes a rotation of each cross section of the circular rod as a whole
about its centre. There is no dispersion for waves of this type, and both the
phase and group velocities are equal to 1.
The dispersion relation


C11 C13


(4.84)
C31 C33 = 0
corresponds to the longitudinal waves, i.e., waves involving the displacements
w1 , w% which are independent of . Expanding (4.84) we obtain
41 q1 (p22 + 2 )Z1 (q1 /2)Z1 (q2 /2) (p22 2 )2 Z0 (q1 /2)Z1 (q2 /2)
41 2 q1 q2 Z1 (q1 /2)Z0 (q2 /2) = 0. (4.85)
Historically, this equation was first derived by Pochhammer in 1876, but
because of its complexity detailed calculations of the dispersion curves did
not appear until much later. The dispersion curves of the first branch of
L-waves according to the one- and three-dimensional theories for = 0.31
are shown in Figure 4.5. Note that the dispersion relation according to
the one-dimensional
theory for the rod of circular cross section, in terms of
p
= h/ / and = kh, reads (cf. (4.65))
2 = 2(1 + )2 .

154

CHAPTER 4. ELASTIC RODS

One can see that the dispersion curves according to the one- and threedimensional theories are practically identical for < 1.2. As the exact dispersion curve approaches asymptotically the straight line = (cr /cs )
from above.

Figure 4.6: Dispersion curves of the lowest branch of F-waves for n = 1: a)


1-D theory: dashed line, and b) 3-D theory: solid line.
The case of n = 1 corresponds to the first family of flexural waves. All
displacement components are not zero, so we have to find the roots of the
equation (4.79). It yields many branches of the dispersion curves in the , plane. However, here we study only the lowest branch in the low-frequency
short-wave region in order to compare with the dispersion curve according
to (4.68). The dispersion curves of the first branch of F-waves in circular
rods according to the one- and three-dimensionalp
theories for n = 1, = 0.31
are shown in Figure 4.6. In terms of = h/ / and = kh the 1-D
dispersion relation reads
1+ 4
2 =
.
8
Again, in the long-wave range ( < 0.6) the difference between them is
negligibly small. However, as the exact dispersion curve approaches
the line = (cr /cs ) from below, which means that the phase velocity is
equal to the Rayleigh wave velocity, in contrast to the unbounded phase
velocity predicted by the 1-D theory.
Problems

4.5. FREQUENCY SPECTRA

155

1. Solve the equation (4.63) subject to the following initial conditions


u(x, 0) = f (x),

u(x,
0) = v(x).

2. Plot the dispersion curves for the flexural waves in an infinite straight
rod. Find the phase and group velocities as functions of the wave
number.
3. Solve the equation of flexural waves (4.67) subject to the initial conditions
u1 (x, 0) = f (x), u 1 (x, 0) = v(x).
4. Show that Pochhammers equation (4.85) yields the following asymptotic formula for the lowest branch of L-waves
2 = 2(1 + )2 + O(4 )
in the low-frequency long-wave range.

5. Show that the dispersion equation (4.79) with n = 1 yields the following
asymptotic formula for the lowest branch of F-waves
2 =

1+ 4
+ O(6 )
8

in the low-frequency long-wave range.


6. Prove that the lowest L- and F-branch of the dispersion curves according to 3-D elasticity approach the line = (cr /cs ) as .

4.5

Frequency spectra

Frequency spectra of straight rods


Longitudinal vibrations. Consider a straight rod of length L. The governing
equation of the longitudinal vibrations is given by (4.63). In order to see
more details of the low frequency spectrum, let us introduce the following
dimensionless variables
s
x
t E
, = .
=
L
L
Equation (4.63) becomes
u| = u00 .

(4.86)

156

CHAPTER 4. ELASTIC RODS

We look for solutions of the form


u(, ) = u()ei ,
where is a dimensionless frequency. Substituting this into (4.86) we obtain
u00 + 2 u = 0,
from which
u = a cos + b sin .
The constants a, b should be determined so as to satisfy the boundary conditions at the ends of the rod. As an example consider now the longitudinal
vibrations of a rod with free ends. In this case the boundary conditions read
u0 (0) = u0 (1) = 0.
These conditions will be satisfied if b = 0 and
sin = 0.
Thus, the eigenfrequencies for the case under consideration are given by
n = n,

n = 0, 1, 2, . . . .

If one end of the rod is fixed and the other is free, the boundary conditions
are
u(0) = u0 (1) = 0.
In this case a = 0 and
cos = 0.
The frequency spectrum is given by
1
n = (n + ),
2

n = 0, 1, . . . .

Torsional vibrations. Introducing the dimensionless variables


s
t
C
x
=
, = ,
L Ip
L

4.5. FREQUENCY SPECTRA

157

we transform the equation of the torsional vibrations (4.66) to


| = 00 .
This equation is identical in form with equation (4.86), and the previous
results can be used in various particular cases.
Flexural vibrations. We assume again that the rod vibrates in one of the
principal planes, so that only u1 6= 0. The governing equation is given by
(4.67). Introducing the dimensionless variables
s
EI11
t
x
= 2
, = ,
L
|S|
L
we transform (4.67) to
u1| + u0000
1 = 0.

(4.87)

We seek the solution of (4.87) in the form


u1 = u1 ei ,
where is the dimensionless frequency. Substituting this formula into (4.87)
we obtain
2
u0000
1 = 0.
1 u

(4.88)

By using the notation

(4.89)

we present the general solution of (4.88) in the form


u1 = a1 (cos + cosh ) + a2 (cos cosh )
+ a3 (sin + sinh ) + a4 (sin sinh ),
in which a1 , . . . , a4 are constants which should be determined from the boundary conditions.
In the case of fixed edges the boundary conditions are
u1 (0) = u001 (0) = 0,

u1 (1) = u001 (1) = 0.

(4.90)

From the first two conditions of (4.90) it can be concluded that a1 = a2 = 0.


From the last two conditions we obtain a3 = a4 and
sin = 0.

158

CHAPTER 4. ELASTIC RODS

The roots of this equation are given by


n = n,

n = 1, 2, . . . ,

and the corresponding frequencies can be determined from (4.89).


If the edges of the rod are free we have the following boundary conditions
u001 (0) = u000
1 (0) = 0,

u001 (1) = u000


1 (1) = 0.

(4.91)

In order to satisfy the first two conditions of (4.91) we have to take a2 =


a4 = 0 so that
u1 = a1 (cos + cosh ) + a3 (sin + sinh ).
From the last two conditions of (4.91) we then obtain
a1 ( cos + cosh ) + a3 ( sin + sinh ) = 0,
a1 (sin + sinh ) + a3 ( cos + cosh ) = 0.

(4.92)

A non-trivial solution of (4.92) exists only if its determinant is equal to zero.


In this manner the following frequency equation is obtained:
cos cosh = 1.

(4.93)

The first four roots of this equation are given below:

Figure 4.7: Modes of the flexural vibrations.


0 = 0,

1 = 4.730,

2 = 7.853,

3 = 10.996.

The corresponding frequencies can be calculated by using (4.89). Substituting the roots of (4.93) into (4.92), the ratios a1 /a3 for the corresponding
modes of vibration and consequently the shape of the vibrating rod can be

4.5. FREQUENCY SPECTRA

159

determined. In Figure 4.7 the shapes corresponding to 1 , 2 and 3 are


shown. The zero root 0 corresponds to the rigid-body motion of the rod.
Frequency spectrum of a closed circular ring. We illustrate the application of the theory of naturally curved rods by considering the free vibrations
of a rod which, in the unstressed state, forms a closed circular ring. We
assume that the cross section of the ring has an axis of symmetry lying in
the plane of the ring. The equation of the central line reads
z = (R cos , R sin , 0),

[0, 2].

Here R is the radius of the central line, x = R is its length. The triad
{t , t} is given by
t1 = ( cos , sin , 0), t2 = (0, 0, 1),
t = ( sin , cos , 0).

(4.94)

It is easy to see that


1
1 = ,
R

2 = $ = 0.

The equations of motion break up into two systems of equations for u, u1 and
u2 , , respectively. The first type of vibrations involving u and u1 is called
plane, the second one involving u2 and flexural-torsional. It may be
shown that the vibrations of a curved rod fall into such classes whenever the
central line of the unstressed rod is a plane curve, and one of the principal
axes lies in its own plane at each point.
Plane vibrations. The equations describing the plane vibrations are as follows
1
1 00
1 0
u1 ) EI11 (u000
u ),
1
R
R
R
1
1
1 000
|S|
u1 = E|S|(u0 u1 ) + EI11 (u0000
u ).
1
R
R
R
|S|
u = E|S|(u00

(4.95)

We introduce the dimensionless quantities


s
t E
x
I11
=
, = , ? =
.
R
R
|S|R2
Equations (4.95) then take the form
00
u| = u00 u01 ? (u000
1 u ),
000
u1| = u0 u1 + ? (u0000
1 u ),

(4.96)

160

CHAPTER 4. ELASTIC RODS

where the prime now denotes the derivative with respect to . Since the ring
is closed, let us seek solutions of (4.96) in the form
(
(
sin n
cos n
i
i
u = ae
, u1 = be
,
(4.97)
cos n
sin n
where a, b are unknown constants to be determined and n is an integer. The
periodic sine and cosine functions of used in (4.97) guarantee that the
displacements are continuous with respect to . Substitution of (4.97) in
(4.96) gives
(n2 + ? n2 2 )a (n + ? n3 )b = 0,
(n + ? n3 )a + (1 + ? n4 2 )b = 0.
The determinantal equation, for ? = 0.0001, yields the following roots
m
1
2

n=0
0.
1.

n=1
0.
1.414

n=2
0.027
2.236

n=3
0.076
3.163

The first two zero roots at n = 0 and n = 1 correspond to the rigid-body


motions of the ring. The second root = 1 at n = 0 describes the radial
vibration of the ring with the frequency much higher than that of = 0.027 at
n = 2, which corresponds to the first fundamental mode of plane vibrations.
Flexural-torsional vibrations. The differential equations describing the vibrations of this type are
1
1
1 00
) + C(00 + u002 ),
R
R
R
1
1
1
Ip = C(00 + u002 ) EI22 (u002 + ).
R
R
R

|S|
u2 = EI22 (u0000
2 +

(4.98)

We introduce the following new quantities and parameters


s
t E
x
=
, = , = R,
R
R
I22
C
Ip
=
,

=
,

=
.
|S|R2
E|S|R2
|S|R2
Equations (4.98) then take the dimensionless form
00
00
00
u2| = (u0000
2 + ) + ( + u2 ),
| = ( 00 + u002 ) (u002 + ).

(4.99)

4.5. FREQUENCY SPECTRA

161

In a similar manner we seek solutions of (4.99) in the form


(
(
sin
n
sin n
u2 = aei
, = bei
.
cos n
cos n
We then obtain the following linear homogeneous equations with respect to
a, b
(n4 + n4 2 )a + ( + )n2 b = 0,
( + )n2 a + ( + n2 2 )b = 0.
The lowest roots of the determinantal equation for = 0.0001, = 0.00012
and = 0.0002 are given in the following table
m
1
2

n=0
0.
0.707

n=1
0.
1.049

n=2
0.027
1.703

n=3
0.077
2.43

Similarly as in the previous case, the first two zero roots correspond to the
rigid-body translation and rotation of the ring. The second root = 0.707
at n = 0 describes the pure torsional vibration, while the first root = 0.027
at n = 2 corresponds to the fundamental mode of vibration with the lowest
frequency.
Problems
1. Find the eigenfrequencies and eigenfunctions for longitudinal vibration
of a straight rod of length L if both ends are fixed.
2. Find the eigenfrequencies and eigenfunctions for flexural vibration of a
straight rod of length L with fixed edges.
3. Solve the preceding problem assuming that one edge is clamped and
the other is free.
4. Study the forced flexural vibrations of a rod with fixed ends under the
force F1 .
5. Find the eigenfrequencies and eigenfunctions for plane and flexuraltorsional vibrations of an open circular ring with free and fixed edges.

162

CHAPTER 4. ELASTIC RODS

Chapter 5
Piezoelectric shells
5.1

Two-dimensional equations

Geometry of a piezoelectric shell and location of electrodes. Let B


denote a domain in the three-dimensional Euclidean space E, which can be
obtained from S, a two-dimensional smooth surface bounded by a smooth
closed curve S, and from the segments of length h in the direction perpendicular to the surface, by the method described in Section 3.1. A linear
piezoelectric body occupying the domain B in its stress-free undeformed state
is called a piezoelectric shell, the surface S its middle surface, and h its thickness. The shell is said to be thin if h is much smaller than the characteristic
sizes as well as the radius of curvature of the middle surface.
The domain B E is specified by the equation
z i (xa ) = ri (x ) + xni (x ),

(5.1)

where z i = ri (x ) is the equation of the middle surface S, and ni (x ) are the


cartesian components of the normal vector n to this surface. The co-ordinates
x take values in a domain S R2 , while x [h/2, h/2].
Let S denote the face surfaces of the shell given by (5.1) for x = h/2.
We consider the three methods of electrode arrangement encountered most
often:
i) There are no electrodes on the face surfaces of the shell (unelectroded
face surfaces). The edge of the shell is partially electroded (see Figure
5.1). Thus, we assume that the contour S is decomposed into open
(1)
(n)
curves ce , . . . , ce (where there are electrodes) and the remaining part
(i)
cd . For xa ce [h/2, h/2] on the electroded portions of the edge
the electric potential is prescribed
= (i) (t),
163

i = 1, . . . , n.

(5.2)

164

CHAPTER 5. PIEZOELECTRIC SHELLS

Figure 5.1: Partially electroded edge of a piezoelectric shell.


On the unelectroded portion of the boundary the electric charge should
vanish.
ii) The face surfaces S are fully coated by the electrodes. They form the
two equipotential surfaces where
= 0 (t)/2,

for x = h/2.

(5.3)

The difference between these values, 0 (t), is sometimes called voltage


for short.
iii) The face surfaces of the shell are only partially coated by electrodes.
This case can be regarded as the mixed situation of the two cases above.
Kinematics of a two-dimensional piezoelectric shell. If we regard
piezoelectric shells as a two-dimensional continuum and try to construct for
them a two-dimensional theory, then it turns out that the number of unknown
functions in such a theory depends on the location of electrodes. In case i)
the kinematics of a piezoelectric shell is completely specified by the two
fields, namely u(x , t) corresponding to the mean displacements of the shell,
and (x , t) being the two-dimensional electric potential. These fields are
assumed to be continuous and as many times differentiable as required. The
measures of extension and bending are defined by (cf. (3.1) and (3.4))
A = u(;) b u,

(5.4)

and
B = u; + (u b( );) + b( u;) c u,

(5.5)

5.1. TWO-DIMENSIONAL EQUATIONS

165

respectively, where u = t u and u = n u are the tangential and normal


components of the displacement vector referred to the basis {t , n}. Here
a , b and c correspond to the first, second and third quadratic forms
of the middle surface, respectively. The raising and lowering of indices of
two-dimensional tensors will be done with the help of the metric tensors
a and a . The semicolon preceding Greek indices denotes co-ordinate
expressions of the covariant derivatives. One can also use the alternative
measures of bending as given by (3.8). The two-dimensional electric
field F is calculated from the electric potential according to
F = , .

(5.6)

In case ii) the kinematics of a piezoelectric shell is specified by the mean


displacement field u(x , t) only, exactly as for the elastic shells. The twodimensional electric enthalpy depends on the measures of extension and bending, referred to as state variables in this case. Finally, in case iii) we must
combine the two types of kinematics described above. In the unelectroded
portion of the face surfaces we need two-field kinematics as in case i), while
in the electroded portion of the face surfaces, the mean displacement field is
enough to specify the kinematics of the shell.
Since u depends on t we introduce the following quantities
v = u (velocity),

(acceleration),
a=u

to measure its time rates. The time rates of do not play any role in the
two-dimensional theory.
Variational principles and boundary-value problems
Case i) For a piezoelectric shell with unelectroded face surfaces the twodimensional action functional reads
Z t1 Z
( ) da dt,
(5.7)
J[u, ] =
t0

where is the kinetic energy density and the electric enthalpy. The action
functional J[u, ] is defined on the space of all functions u and , where u is
assumed to be continuously twice differentiable, and the remaining functions
are continuously differentiable. The kinetic energy density is a quadratic
form of u
1
1
u = (u 2 + u u ),
(5.8)
= u
2
2

166

CHAPTER 5. PIEZOELECTRIC SHELLS

where is the mass density per unit area. The electric enthalpy is a quadratic
form of the measures of extension and bending and of the electric field, which,
in the general case, is given by
0 (A , B , F ),
=
or
0 (A , , F ).
=
If the edge of the shell is free, then it is natural to assume that no constraints are imposed on u at the boundary. If the edge of the shell is clamped,
we assume that J[u, ] is defined on the space of admissible displacement
fields u satisfying the boundary condition
u = 0,

u = 0,

u, = 0 at Sk ,

(5.9)

where denotes the surface vector normal to the curve S. The last of
these constraints expresses the fact that the rotation angle of the edge of the
shell vanishes (clamped edge). If, finally, the edge of the shell is fixed, then
only the displacements at Sk should vanish
u = 0,

u = 0,

at Sk .

(5.10)

Concerning the admissible continuously differentiable 2-D electric potentials


, we require that they satisfy the following boundary conditions
= (i) (t) on c(i)
e , i = 1, . . . , n.

(5.11)

The conditions (5.11) can be interpreted as the conditions (5.2) on average.


They agree with the generalization of Saint-Venants principle to piezoelectric
shells [50].
We also require that u is given at t = t0 and t = t1
u|t=t0 = u0 ,

u|t=t1 = u1 .

Hamiltons variational principle for piezoelectric shells states that the true
and electric potential correspond to the stationary
displacement field u
points of the action functional (5.7)
J = 0.
In order to derive the equations of motion for piezoelectric shells let us
calculate the variation of the functional (5.7)
Z t1 Z
J =
[
(u
u + u u ) N A M B + G F ] da dt, (5.12)
t0

5.1. TWO-DIMENSIONAL EQUATIONS

167

where the tensors N , M and G are defined according to


0

,
A
0

M =
,
B
0

G =
.
F
N =

(5.13)

We call N (symmetric) membrane stresses, M bending moments and G


2-D electric induction field.
According to (5.4)-(5.6) the variations of A , B and F are equal to
A = u(;) b u,
B = u; + (u b( );) + b( u;) c u,
F = , .
We substitute these formulae for the variations into (5.12). Assuming the
regularity of all the quantities under the integral sign, we transform the latter
with the help of Gauss theorem for the variations vanishing at the boundary
S to
Z t1 Z
J =
[(
u + T; + b M; )u + (
u + T b
t0

M;
)u + G; ] da dt = 0,

where T is the unsymmetric tensor given by


T = N + b M .
Since u and are arbitrary inside the region S (t0 , t1 ), we conclude that
u = T; + b M; ,

u = T b M;
,

(5.14)

G; = 0.
These are the two-dimensional equations of motion of the piezoelectric shell.
Substituting (5.13) into (5.14), we obtain four differential equations with
respect to the four unknown functions, namely, the three components of the
displacement field u and the electric potential .

168

CHAPTER 5. PIEZOELECTRIC SHELLS

For the variations not vanishing at the boundary we can repeat the procedure similar to that in Section 3.1 to obtain
Z t1 Z

{(T b M ) u + [M; + (M )]u


s
t0
Ss
M u; G } da dt = 0.
For the free edge the variations u, u and u; are arbitrary at S, while
is arbitrary only on cd ; hence
T + b M = 0,

M;
+ (M ) = 0,
s

M = 0,

(5.15)

= (i) (t) on c(i)


e , i = 1, . . . , n,

G = 0 on cd .
The last equation says that the average electric charge vanishes on cd . For
the clamped edge the conditions (5.9) and (5.15)4,5 should be posed at the
boundary. If the edge is fixed, (5.10) and (5.15)4,5 are the boundary conditions at S.
is given in terms of the measures A , and F , the equations of
If
motion read

u = t
, + b m; + F ,

u = n b m
; + F,
G;

(5.16)

= 0,

where
0

,
A
0

m =
,

G =
,
F
n =

and the unsymmetric tensor t is equal to


[

t = n + b m] .

(5.17)

5.1. TWO-DIMENSIONAL EQUATIONS

169

In a similar manner, the following free-edge boundary conditions are obtained


t + b m = 0,

m
(m ) = 0,
; +
s
m = 0.
= (i) (t) on

c(i)
,i

(5.18)

= 1, . . . , n,

G = 0 on cd .
Case ii) For a piezoelectric shell with fully electroded face surfaces the twodimensional action functional depends only on the mean displacement field
u
Z Z
t1

( ) da dt,

J[u] =
t0

(5.19)

1 (A , B )
where is the kinetic energy density given by (5.8) and =
is the electric enthalpy depending only on the measures of extension and
bending (cf. the analogous formulae for elastic shells). The true displacement
is the stationary point of the functional (5.19). Thus, this variational
field u
principle leads to the same equations of motion and boundary conditions as
in the theory of elastic shells (cf. equations (3.19)-(3.21) or (3.24)-(3.26)).
Changes concern just the constitutive equations.
Case iii) Let S0 be a portion of the middle surface, the face surfaces of which
are not electroded, and S1 the remaining part. Assume, for definiteness,
that the latter lies strictly inside S and S1 is the boundary between these
two regions. The 2-D action functional can be presented as a sum of two
integrals
Z t1 Z
0 (A , B , F )] da dt
J[u, ] =
[
t0
S
(5.20)
Z t10Z
1 (A , B )] da dt.
+
[
t0

S1

We assume that the space of admissible displacement fields and electric potentials, on which the functional (5.20) is defined, contains also the displacement fields with discontinuous first and second derivatives at the boundary
S1 . However we require that the displacements u , u as well as the first
derivative u; should be continuous there, where denotes the outward unit
normal to the curve S1 . Concerning the 2-D admissible electric potentials
we require that they vanish at the boundary S1 . The true displacement field

170

CHAPTER 5. PIEZOELECTRIC SHELLS

and electric potential correspond to the stationary points of the functional


(5.20). Varying this action functional with the vanishing variations at the
boundary S1 , we obtain the equations of motion (5.14) in the region S0 and
(3.19) and (3.20) in S1 . The boundary conditions at the free edge of the shell
S are given by (5.15).
For the variations not vanishing at the boundary S1 we obtain the following equation
Z t1 Z

{[[T b M ]] u + [[M; + (M )]]u


s
t0
S1
[[M ]] u; } da dt = 0,

(5.21)

where [[.]] denotes the jump of the corresponding quantity across the line
S1 : [[A]] = A+ A , the indices + and indicating the limiting values of
A as x approach the two sides of the line S1 . Due to the arbitrariness of
u , u and u; equation (5.21) yields the following jump conditions at the
boundary line S1 of the unelectroded region
[[T + b M ]] = 0,

[[M;
+ (M )]] = 0,
s
[[M ]] = 0.

(5.22)

Since the electric potential vanishes at S1 there is no jump condition for


the electric induction field.
2-D electric enthalpy and constitutive equations
Case i) Assume that the shell is made of a linear homogeneous piezoelectric
material. It turns out that, within the first approximation, the 2-D electric
enthalpy density is given by
2
0 = h (c A A + h c B B

2 N
12 N

2e
N A F N F F ),

(5.23)

where c
, e
N
N , N are the so-called two-dimensional electroelastic moduli.
On the basis of (5.13) and (5.23), the following relations may be established

A e
N = h(c
N F ),
N
h3
M = c
B ,
12 N

G = h(e
N A + N F ).

(5.24)

5.1. TWO-DIMENSIONAL EQUATIONS

171

These are the constitutive equations for the two-dimensional theory of piezoelectric shells with unelectroded face surfaces.
Case ii) If the face surfaces of the piezoelectric shell are fully electroded, the
electric enthalpy depends only on the measures of extension and bending and
is given by
2
1 = h (c A A + h c B B + 2e3 A 0 (t) ).

P
2 P
12 N
h

(5.25)

The constitutive equations are similar to those for elastic anisotropic shells
and take the form
N = h(c
A + e3
P
P
M =

h3
c
B .
12 N

0 (t)
),
h
(5.26)

For the calculation of the mechanical eigenfrequencies under short-circuit


conditions 0 (t) must be set equal to zero.
Case iii) In the region S0 the electric enthalpy is given by (5.23), in the
remaining region S1 by (5.25). The corresponding constitutive equations
take the form (5.24) and (5.26), respectively.
One can easily obtain the alternative constitutive equations, if the measure of bending enters the electric enthalpy instead of B . In exactly
the same manner as in Section 3.1, one can prove the asymptotic equivalence
of the derived constitutive equations.
Problems
1. Show that the equations of motion and the boundary conditions for
piezoelectric plates with b = 0 break up into those of longitudinal
and flexural vibrations. Find the corresponding uncoupled equations
and boundary conditions in terms of u and u.
2. Derive the balance equation of energy for piezoelectric shells. Using
this equation, prove the uniqueness of solutions of the boundary-value
problems.
3. Formulate the variational principles for the eigenvalue problems of
determining the resonant and antiresonant vibrations of piezoelectric
shells.

172

5.2

CHAPTER 5. PIEZOELECTRIC SHELLS

Asymptotic analysis

Three-dimensional action functional. The asymptotic analysis of the


action functional of piezoelectric shells is technically more cumbersome than
that of elastic shells due to the presence of the electric potential in the action
functional and due to the fact that the 3-D electric enthalpy, in the general
case of anisotropy, contains various additional terms. To ease the subsequent
asymptotic analysis we want first to reformulate the exact variational problem for piezoelectric shells. As before we refer a piezoelectric shell to the
curvilinear co-ordinates x , x in the domain B E specified by
z i (xa ) = ri (x ) + xni (x ),
where z i = ri (x ) is the equation of the middle surface S bounded by the
contour S, and ni (x ) denote the cartesian components of the normal vector n to this surface. The co-ordinates x take values in the domain S R2 ,
while x [h/2, h/2]. The co- and contravariant components of the metric
tensor in the curvilinear co-ordinate system {x , x} are given by the formulae (3.33) and (3.36). Defining the characteristic radius of curvature R as
described in Section 3.2, we assume that
h
 1.
R
Concerning the location of electrodes we restrict ourselves to the three cases
i)-iii) described in the previous section. Since case iii) is simply the combination of the first two cases, the asymptotic analysis will be done for the first
two cases only.
According to the variational principle (2.45), (2.47) the true displacement
field wi and the electric potential correspond to stationary points of the
action functional
Z t1 Z Z h/2
I=
(T W ) dx da dt
(5.27)
t0

h/2

under the constraints (5.2) in case i), and the constraints (5.3) in case ii),
where = 1 2Hx + Kx2 and da denotes the area element. In (5.27) T is
the kinetic energy density
1
1
T = w i w i = (a w w + w 2 ),
2
2
with w , w being the projections of the displacement vector onto the tangential and normal directions to the middle surface
w = ti wi ,

w = ni wi .

5.2. ASYMPTOTIC ANALYSIS

173

The electric enthalpy density W is the quadratic form of the strain ab and
the electric field Ea and is given by
1
1
cab
ab Ec ab
Ea Eb .
W (ab , Ea ) = cabcd
E ab cd e
2
2 S

(5.28)

The problem is to replace the three-dimensional action functional (5.27)


by an approximate two-dimensional functional for a thin shell, whose functions depend only on the longitudinal co-ordinates x1 , x2 and time t. The
possibility of reduction of the three- to the two-dimensional problem is related
to the smallness of the ratios between the thickness h and the characteristic radius of curvature R of the shell middle surface and between h and the
characteristic scale of change of the electroelastic state in the longitudinal
directions l. Additionally, we assume that
h
 1,
c

(5.29)

where is the characteristic scale of change of the function wi in time (see


Section 2.5) and c is the minimal velocity of plane waves in the piezoelectric
material under consideration. This means that we consider in this Chapter only low-frequency vibrations of the piezoelectric shell. By using the
variational-asymptotic method, two-dimensional action functionals will be
constructed below in which terms of the order h/R and h/l are neglected as
compared with unity (the first-order or classical approximation).
In order to fix the domain of the transverse co-ordinate in the passage to
the limit h 0, we introduce the dimensionless co-ordinate
=

x
,
h

[1/2, 1/2].

Now h enters the action functional explicitly through the components of the
strain tensor ab and the electric field Ea (cf. (3.40))
 = w(;) b w hb( w;) + hc w,
1
23 = w| + w, + b w b w| ,
h
1
33 = w| ,
h
1
E = , , E3 = | .
h

(5.30)

Here the vertical bar followed by indicates the partial derivative with respect to and not with respect to x . We denote by h.i the integral over
within the limits [1/2, 1/2].

174

CHAPTER 5. PIEZOELECTRIC SHELLS

Another form of the electric enthalpy. Before applying the variationalasymptotic procedure let us transform the electric enthalpy density to another form more convenient for the asymptotic analysis. We note that among
terms of W (ab , Ea ) the derivatives w, /h and w, /h in 3 and 33 as well
as E3 are the main ones in the asymptotic sense. Therefore it is convenient
to single out the components 3 and 33 as well as E3 in the electric enthalpy. We represent the electric enthalpy density W (ab , Ea ) as the sum of
two quadratic forms Wk and W corresponding to longitudinal and transverse
electric enthalpies, respectively. These are defined by
Wk = min max W,
3 ,33 , E3

W = W Wk .
Long, but otherwise simple calculations show that
1
1
  e
Wk = c
N
N  E N E E ,
2
2
1 3333 2
1
+ c33

W = cE + c333
E
2
2 E
1
F 2,
e333 F e33 F 33
2 S

(5.31)

where
= 33 + r  r E ,

= 23 + t
 t E ,
F = E3 + q  + q E .
33 333 3333 333 33 33
The coefficients c
, e
,cE ,cE ,e ,e ,S ,r ,r ,t , t , q
N
N ,N ,cE

and q can be regarded as components of surface tensors referred to the


basis vectors t of the middle surface. We shall call them two-dimensional
electroelastic moduli. They are evaluated in terms of the three-dimensional
moduli by means of the formulae

= c
+ q e3
c
P
P ,
N

N = P q P ,

= e
q 3
e
P
P ,
N

33
q = e3
P /P ,

c
= c k c3 ,
P
b + k eb3 ,
b
P =

33
q = 3
P /P ,

ea
= ea k ea3 ,
P
33
33 + k e33 ,
P =

(5.32)

5.2. ASYMPTOTIC ANALYSIS


k = h c3 ,

175

3333
cab = cab
ca33
cb33
E
E
E /cE ,
a33 b33 3333
ab = ab
e /cE ,
S +e

t = k k q ,
f =

h = (
c33 )1 ,

k = h e3 , k = h e33 ,

c33
c333
E k
E
,
c3333
E

a33 3333
eab = eab cb33
/cE ,
E e

t
= k + k q ,

r = f + f q ,
f =

r = f f q ,

e33 c333
E k
,
c3333
E

f=

e333 c333
E k
.
3333
cE

Two-dimensional tensors of electroelastic moduli. From (5.32) one


can see that the two-dimensional electroelastic moduli c
, e
and 
N
N
N
satisfy the following symmetry properties
c
= c
= c
= c
,
N
N
N
N
e
= e
N
N ,

N = N .

The same can be said about the 2-D moduli c


and e3
P
P .
We now consider piezoelectric shells that are homogeneous over the thickness. The components of the 3-D moduli of such a shell, referred to the basis
{e , e}, should not depend on . But if these tensors are referred to the basis
{t , n}, their components will in general depend on through the shifter ,
as it is seen from the relation (3.32). Therefore the components of the 2-D
moduli depend in general also on . It can be shown for the homogeneous
shell, however, that components of two-dimensional electroelastic moduli of
any rank, denoted symbolically by A(x , ), possess the property
h
A(x , ) = A0 (x ) + O( )A0 (x ),
R
where A0 are their values evaluated at = 0 and the term O(h/R)A0 is due to
the shifter solely. Therefore, when constructing 2-D shell theories having
the error h/R as compared with unity, it can be assumed that A = A0 , i.e.,
the 2-D moduli of the shell homogeneous over the thickness are independent
of the transverse co-ordinate .
Let us note certain special symmetry cases.
Mirror planes parallel to the middle surface. If properties of the piezoelectric material are invariant under reflections relative to planes parallel to the middle surface, then the following 2-D tensors vanish
c333
= 0, e333 = 0,
E

t
= 0, t = 0,

q = 0, q = 0,

and
c
= c
,
N
P

e
= e
N
P ,

N = P .

176

CHAPTER 5. PIEZOELECTRIC SHELLS

n-fold rotation axes that coincide with the normal to the middle surface.
When n is even, all 2-D tensors of odd rank vanish

= 0, e33 = 0, r = 0, t
e
= 0, c333
= 0, q = 0.
E
N

Transverse isotropy. When properties of the piezoelectric material


are invariant under rotations about the normal to the middle surface
(model of a piezoceramic shell polarized along the normal with symmetry m), it can be shown that all 2-D tensors of odd rank vanish;
the tensor c
has the form
N


+ a a ),
+ cN
c
= cN
2 (a a
1 a a
N

and all the 2-D tensors of second rank are spherical.


Asymptotic analysis of the action functional
Case i) Unelectroded face surfaces. We could start the variational-asymptotic procedure with the determination of the set N according to the general
scheme as it was done for elastic shells (see Section 3.2). As a result, it would
turn out that, at the first step, the functions w and do not depend on the
transverse co-ordinate : w = u(x , t), = (x , t); at the second step the
function w? is a linear function of ; and at the next step w?? and ?? are
completely determined through u and . Thus, the set N according to the
variational-asymptotic scheme consists of functions u(x , t) and (x , t). We
will pass over these long, but otherwise standard, deliberations and make a
change of unknown functions immediately.
We introduce the following functions
u (x , t) = hw (x , , t)i,

u(x , t) = hw(x , , t)i,


(x , t) = h(x , , t)i.

(5.33)

The functions u , u correspond to the mean displacements of the shell, while


describes the mean electric potential. Now let us make the following change
of unknown functions (cf. (3.45))
w = u (x , t) h (x , t) + hy (x , , t),
w = u(x , t) + hy(x , , t),
= (x , t) + h(x , , t),

(5.34)

where
= u, + b u .

(5.35)

5.2. ASYMPTOTIC ANALYSIS

177

Because of the definitions (5.33) the functions y , y and should satisfy the
following constraints
hy i = 0,

hyi = 0,
hi = 0.

(5.36)

Equations (5.34) and (5.36) set up a one-to-one correspondence between


w , w, and the set of functions u , u, , y , y, and determine the change
in the unknown functions {w , w, } {u , u, , y , y, }.
Asymptotic analysis enables us to determine the order of smallness of
y , y, . If these terms are neglected, then (5.34) is a generalization of the
well-known Kirchhoff-Love hypotheses to a piezoelectric shell. The electroelastic state of a shell is then characterized by the measures of extension A ,
the measures of bending B and, finally, the surface electric field F = , ,
where A and B are given by (5.4) and (5.5), respectively. We introduce
the following notation
q
q
p
A = max A A , B = h max B B , fF = max F F ,
S

= max |y| |,
B

= max |y| |,
B

= max || |.
B

Consider a certain point of the middle surface S. The best constant l in the
inequalities
|A, |
max |y, |

A
,
l

,
l

B
fF
, |F, |
,
l
l

max |y, | , max |, |

l
l
h |B, |

is called the characteristic scale of change of the electroelastic state in the


longitudinal directions. We define the inner domain Si as a subdomain of S
in which the following inequalities hold:
h? = h/R  1,

h?? = h/l  1.

(5.37)

We assume the domain S to consist of the inner domain Si and a domain


Sb abutting on the contour S with width of the order h (boundary layer).
Then the functional (5.27) can be decomposed into the sum of two functionals, an inner one for which an iteration process will be applied, and a
boundary layer functional. As in the theory of elastic shells, the boundary
layer functional can be neglected in the first-order approximation. Therefore,
the problem reduces to finding stationary points of the inner functional that
can be identified with the functional (5.27) (Si S).

178

CHAPTER 5. PIEZOELECTRIC SHELLS

We now fix u , u, and seek y , y, . Substituting (5.34) into the action


functional (5.27) and taking (5.29) into account, we neglect the time rates of
y , y in the kinetic energy. We now estimate terms in the electric enthalpy.
Repeating the estimations based on the inequalities (5.37) similar to those
provided in (3.48)-(3.50) we can obtain the asymptotic formulae
 = A hB ,

23 = y| ,

33 = y| .

(5.38)

It is also easy to check that, within the first-order approximation,


E = F ,

E3 = | .

(5.39)

According to the equations (5.38) and (5.39) the partial derivatives of y , y,


with respect to x do not enter the action functional. The functions y , y,
do not enter the longitudinal electric enthalpy. Putting 1 we obtain the
following functional
h
I =
2

t1

t0

2
333
hc3333
+ c33

E + 2cE
E

S
2
2e333 F 2e33 F 33
S F i da dt, (5.40)

where
= y| + r (A hB ) r F ,

= y| + t
(A hB ) t F ,
F = | + q (A hB ) + q F .
We maximize the functional (5.40) in y , y and minimize in under the
constraints (5.36). The minimax value of I is equal to zero and is attained
at = = F = 0, i.e., at
1
1
y = (r A r F ) + hr B ( 2 ),
2
12
1
1

2
y = (t
),
A t F ) + ht B (
2
12
1
1
= (q A + q F ) hq B ( 2 ).
2
12

(5.41)

Regarding (5.34) as the asymptotic Ansatz, with u , u and the unknown functions, and with y , y and given by (5.41), we can substitute it
into the action functional (5.27) and integrate over the thickness to obtain

5.2. ASYMPTOTIC ANALYSIS

179

the average Lagrangian. On the fields (5.34) the average transverse electric enthalpy vanishes, while the principal terms of the average longitudinal
electric enthalpy give
0 = h hc (A hB )(A hB )

2 N

2e
N (A hB )F N F F i
h
h2

= (c
A
A
+
c
B B 2e

N A F N F F ).
2 N
12 N
This is exactly the formula (5.23). Since the kinetic energy density contains
only w and w,
whose principal terms are the same as for elastic shells, the
average two-dimensional kinetic energy is found to be
1
= h(u 2 + a u u ).
2
Thus, the two-dimensional action functional (5.7) is justified for this case.
Case ii) Fully electroded face surfaces. In this problem the electric potential
should satisfy the constraints (5.3). Consequently, we make another change
of the unknown functions:
w = u (x , t) h (x , t) + hy (x , , t),
w = u(x , t) + hy(x , , t),
= 0 (t) + h(x , , t),

(5.42)

where is given by (5.35). Thus, the difference between (5.34) and (5.42)
concerns only the first term of , where 0 is substituted in place of . We
impose the constraints
hy i = 0, hyi = 0,
|=1/2 = 0

(5.43)

on the functions y , y, .
Let us introduce the following notation
q
q
A = max A A , B = h max B B ,
S

= max |y| |,
B

= max |y| |,
B

= max || |.
B

We define the characteristic scale of change of the electroelastic state in the


longitudinal directions as the best constant l in the inequalities
A
B
|A, | , h |B, | ,
l
l

max |y, |
, max |y, | , max |, | ,

l
l
l

180

CHAPTER 5. PIEZOELECTRIC SHELLS

and we make the same assumption as in (5.37). By performing an estimation


procedure analogous to the previous case, it can be shown that the following
asymptotic formulae
 = A hB ,

23 = y| ,

E = 0,

33 = y| ,
0
E3 = |
h

(5.44)

hold true within the first-order approximation. Fixing u , u and substituting


(5.42) into the action functional (5.27), we determine y , y, as the stationary
point of the functional obtained. It can be shown that the kinetic energy and
the longitudinal enthalpy do not contain principal terms involving y , y, .
Keeping the principal terms in the transverse electric enthalpy, we obtain the
functional
Z Z
h t1
hc3333 2 + 2c333
+ c33

I =
E
E
2 t0 S E
2
2e333 F 2e33 F 33
(5.45)
S F i da dt,
where
= y| + r (A hB ),
= y| + t
(A hB ),
0
F = | + q (A hB ).
h
We maximize the functional (5.45) in y , y and minimize it in under the
constraints (5.43). Varying the functional (5.45) we obtain the equations
333
(c3333
e333 F )| = 0,
E + cE

(c333
+ c33
e33 F )| = 0,
E
E
(e333 + e33 + 33
S F )| = 0,

(5.46)

and the boundary conditions at = 1/2


333
e333 F = 0,
c3333
E + cE

c333
+ c33
e33 F = 0,
E
E
= 0.

(5.47)

Equations (5.46) and (5.47) yield


333
c3333
e333 F = 0,
E + cE

c333
+ c33
e33 F = 0,
E
E
e333 + e33 + 33
S F = A,

(5.48)

5.2. ASYMPTOTIC ANALYSIS

181

where A is a constant. Solving (5.48) we obtain


F = C,

= k C,

= f C,

where the constant C should be chosen from the boundary conditions = 0


at = 1/2. From here it follows that C = 0 /h + q A , consequently
F = 0 /h + q A ,

= k F,

= kF,

and
1
0
f A + hr B ( 2
h
2
0
1
y = k k A + ht
B ( 2
h
2
1
1
= hq B ( 2 ).
2
4
y = f

1
),
12
1
),
12

(5.49)

We now regard (5.42) as the asymptotic Ansatz, with u , u the unknown


functions, and with y , y and given by (5.49). Substituting it into the
action functional (5.27) and integrating over the thickness, we obtain the
average Lagrangian. On the field (5.42) E = 0 and the principal terms of
the average longitudinal electric enthalpy are
k = h hc (A hB )(A hB )i

2 N
h
h2
= (cN A A + cN B B ).
2
12
The average transverse electric enthalpy is equal to
h 33
2

2
= h 33

P F = P (0 /h + q A ) .
2
2
Taking the sum of these electric enthalpies we obtain the average electric
enthalpy in the form (5.25). By the same arguments one can show that the
kinetic energy density is given by (5.8), so that the two-dimensional action
functional (5.19) is fully justified.
Relationship between 3-D and 2-D electroelastic states. To complete
the 2-D theory of piezoelectric shells we should also indicate the method
of restoring the 3-D electroelastic state by means of the 2-D one. To do
this, the strain and the electric field E should be found within the firstorder approximation by the asymptotic formulae (5.38), (5.39) or (5.44). The

182

CHAPTER 5. PIEZOELECTRIC SHELLS

stress tensor and the electric induction D are then determined by the 3-D
constitutive equations.
Case i) The 3-D displacements and the 3-D electric potential are restored
from u , u, according to (5.34) and (5.41). Using (5.38) and (5.39) we find
the strain and the electric fields
 = A hB ,
23 = (r A t F ) + ht
B ,
33 = (r A r F ) + hr B ,
E3 = (q A + q F ) + hq B .

E = F ,

Knowing (, E), one can calculate (, D) according to the 3-D constitutive


equations (2.38). While doing so, it is convenient to use the decomposition
(5.31) for the electric enthalpy. Within the first-order approximation we find

= c
A e
B h =
N
N F cN

N 12
2 M ,
h
h

3 = 0,

D = e
N A

33 = 0,
G

+
F

e
B
h
=
e

N
N
N B h,
h

D3 = 0.

All of these equalities should be understood in the asymptotic sense, i.e., they
are accurate up to terms of the orders h/R and h/l of smallness compared
with unity.
Case ii) Calculations are similar to those given in case i). Omitting them,
we present the final formulae for (, E) and (, D).
Strain electric field
 = A hB ,
0
23 = k
k A + ht
B ,
h
0
f A + hr B ,
33 = f
h
0
E = 0, E3 = + hq B .
h
Stress electric induction

c
A
P

3
D = e
P A N

0
e3
P

c
B h
N

h
3 = 0,

N 12
=
2 M ,
h
h

33 = 0,

0
e
N B h,
h

D3 = 33
P

(5.50)
0
+ e3
P A .
h

5.3. ERROR ESTIMATION AND COMPARISON

183

Again, these formulae are accurate up to terms of the orders h/R and h/l of
smallness.
Problems
1. Show that the 3-D equation of electrostatics for a shell in the coordinates x , x reads
(D ); + (D3 ),x = 0.
2. Check the additive decomposition of the electric enthalpy (5.31) and
obtain the formulae (5.32) for the 2-D electroelastic moduli.
3. Follow the variational-asymptotic scheme to determine the set N and
to justify the change of unknown functions (5.34) in the problem i).
4. Perform the similar procedure to justify (5.42) in the problem ii).

5.3

Error estimation and comparison

In this section we shall prove an identity that generalizes Prager-Synges


identity to the statics of piezoelectric bodies. Based on this identity an error
estimate of the piezoelectric shell theory will be established. In some special
cases we shall compare the 2-D constitutive equations obtained in this book
with other well-known equations available in the literature.
Generalization of Prager-Synges identity. Consider the linear vector
space of electroelastic states that consists of elements of the form = (, E),
where is the stress field and E is the electric field; both defined in the
domain B of the three-dimensional piezoelectric body. In this space we introduce the following norm:
Z
2
k kL2 = C2 [] =
G(, E) dv,
(5.51)
B

where the function G(, E) is the density of the complementary energy introduced in Section 2.4. In component form G(, E) reads
1 T a b
1
ab cd
ab c
G(, E) = sE
abcd + dcab E + ab E E .
2
2
Since the complementary energy density G(, E) is positive definite (see
Section 2.4), the definition (5.51) is meaningful.

184

CHAPTER 5. PIEZOELECTRIC SHELLS

for which
We call kinematically admissible those electroelastic states
exist such that
and the electric induction field D
the compatible strain field
1
+ (w)
T ), w
= 0 on Sk ,
= (w

2
= 0, Dn
= 0 on Sd ,
divD
are expressed in terms of
according to equations
and E
and D
while
statically admissible, when
(2.48). We call those electroelastic states
= 0,
div
= ,
E

= 0 on Ss ,
n

= (i)

on Se(i) , i = 1, . . . , n.

= (,
be the true electroelastic state that is realized in a piezo E)
Let
electric body B on the given values of the electric potential (i) on the elec(i)
trodes Se , i = 1, . . . , n. Then the following identity
1 (
+ )]
= C2 [ 1 (
)]

C2 [
2
2

(5.52)

turns out to be valid for arbitrary kinematically and statically admissible


and .
This identity generalizes the well-known Prager-Synge idenfields
)
may be
tity [46] to the statics of piezoelectric bodies. It implies that 12 (+
regarded as an approximation to the true solution, if the complementary
)
is small. In this case we
energy associated with the difference 21 (
or
as an approximation, in view
may also consider each of the fields
of the inequalities
)]
C 2 [
],

C 2 [
)]
C 2 [
],

C 2 [
which follow easily from (5.52).
To prove the identity (5.52) we first rewrite its left-hand side as follows
1 (
+ )]
= C 2 [

1 (
)]

C2 [
2
2
]
+ C2 [ 1 (
)]
[
,

]

= C 2 [
2
1
,

],

= C2 [ (
)] + [
(5.53)
2
where [, 0 ] denotes the scalar product of two elements
Z
Z
0
0
[, ] =
G(, ) dv =
sAB A 0B dv.
B

(5.54)

5.3. ERROR ESTIMATION AND COMPARISON

185

In (5.54) A = ( n , E b ) and sAB is a symmetric matrix, whose elements are


themselves matrices
1
  mn
 E
cD
(h)T mb
smn (d)Tmb
,
(5.55)
=
sAB =
han
Sab
dan Tab
where (d)Tmb = dbm are elements of the transpose matrix and the abbreviated
indicial notation is used. According to (5.53), the identity (5.52) holds true,
when
,

]
= 0.
[

and
and the formulae
This identity follows from the definitions of ,
(5.54) and (5.55). Indeed
Z

E)(
D
D)]
dv
):(

) + (E
[ , ] = [(
ZB
D)]
dv,
):(

w)
+ ( )(
w
D
= [(
B


and (5.55). Integrating
which is the consequence of the definitions of ,
as well as the boundary
this identity by parts and taking the definition of
conditions into account, we can realize that the right-hand side vanishes.
Thus, the identity (5.52) is proved.
Error estimates. From the identity (5.52) the following error estimate can
be established.
Theorem. The electroelastic state determined by the 2-D theory of piezoelectric shells differs in the norm L2 from the exact electroelastic state determined by the 3-D theory of piezoelectricity by a quantity of the order
h/R + h/l as compared with unity.
To prove this theorem it is enough to find out the kinematically and
statically admissible 3-D fields of electroelastic states that differ from the
electroelastic state determined by the 2-D theory by a quantity of the order
h/R + h/l as compared with unity. We construct these fields for the two
cases considered above.
Case i) Construction of kinematically admissible field. We specify the kinematically admissible displacement field in the form
w = u (x ) x (x ) + hy (x , x),
w = u(x ) + hy(x , x),

186

CHAPTER 5. PIEZOELECTRIC SHELLS

where and y , y are given by (5.35) and (5.41), respectively. Here and
below, all quantities without the superscriptsandrefer to the solution of the
equilibrium equations of piezoelectric shells obtained by the two-dimensional
theory. The components of the strain tensor are calculated according to
(5.30). Assume that the 2-D electroelastic state is characterized by the strain
amplitude  = A +B , and the quantity fF defined in Section 5.2 is expressed
through  by fF = c, with c a constant. The asymptotic analysis similar to
that given in (3.48)-(3.50) shows that
 = A xB + O(h/R, h/l) =  + O(h/R, h/l),

23 = r A t F + t
B x + hy, + b hy
= 23 + O(h/R, h/l),

33 = hy,x = (r A r F ) + r B x = 33 .


of the electric induction in the form
We take the components D
= z (x ) e B x.
D
N
We choose z (x ) such that
ix = e A + F =
hD
N
N

G
,
h

(5.56)

where h.ix denotes the integral over x within the limits [h/2, h/2]. The
3 is found by solving the 3-D equation of electrostatics
component D
); + (D
3 ),x = 0
(D

(5.57)

3 = 0 at x = h/2. Due to the choice


subject to the boundary conditions D
the equation (5.57) yields an unique solution
(5.56) for D
Z
1 x
3

D =
(D ()()); d = O(h/R, h/l)
h/2
that satisfies the above-mentioned boundary conditions at x = h/2 (the
3 = 0 at x = h/2 is fulfilled because of the 2-D equation of
condition D
does not satisfy the
electrostatics G; = 0). Note that the constructed field D

= 0, posed at the portion Sd [h/2, h/2]


exact boundary condition D
of the edge, but satisfies it only on average, i.e.
ix =
hD

G
= 0.
h

5.3. ERROR ESTIMATION AND COMPARISON

187

For simplicity of the proof we further assume that the 3-D boundary conditions at the edge of the shell agree with the inner expansion of the electroelastic state (the so-called regular boundary conditions). Then the electric
constructed above is kinematically admissible. For irreguinduction field D
lar boundary conditions we have to take into account an additional electric
induction field that differs substantially from zero only in a thin boundary
layer at the shell edge. Since the energy of this boundary layer is of the order
h/l compared with that of the inner domain, one can easily generalize the
proof of the theorem to this case.
we find
= (,
from the constitutive equations
E)
Knowing (
, D),
= (, D)+O(h/R, h/l), it is easily seen that (,
=
E)
(2.48). Because (
, D)
(, E) + O(h/R, h/l).
Construction of statically admissible field. We write down the 3-D equilibrium equations for a shell in the form (cf. Exercise 2, Section 3.2)
; + ( ),x b = 0,
; + b + ,x = 0,

(5.58)

where
=
,

=
3 ,

=
33 .

Note that is unsymmetric. To find the statically admissible stress field


satisfying (5.58) and the boundary conditions


3 = 0,

33 = 0 at x = h/2,

(5.59)

we proceed as follows. We specify


in the form

= s
0 (x ) xs1 (x ),

where s
0 and s1 are symmetric and independent of x. These are chosen
from the conditions

h
ix = T ,

h
xix = M .

(5.60)

and
The conditions (5.60) enable one to determine s
0 and s1 through T

M uniquely. Moreover, one can check that

s
0
s
1

N
+ O(h/R, h/l),
=
h
12
= 3 M + O(h/R, h/l).
h

188

CHAPTER 5. PIEZOELECTRIC SHELLS

Solving (5.58),(5.59) with the given , we can find and and then

3 and
33 . It turns out that (5.60) are the sufficient conditions for the
existence of and . Indeed, integrating (5.58) and (5.58)1 multiplied by x
over x [h/2, h/2], we obtain
h/2

T; b N + ( )|h/2 = 0,
h/2

N;
+ b T + |h/2 = 0,

(5.61)

h/2

M; N + (x )|h/2 = 0,
where N = h
ix . From the first and the last equations of (5.61) it follows
h/2
that ( )|h/2 = 0, since T; + b M; = 0 according to the 2-D equations
h/2

of equilibrium. From the second equation of (5.61) we also obtain |h/2 = 0.


Thus, if the boundary conditions (5.59) are satisfied at x = h/2, then after
the integration they will also be satisfied at x = h/2. Not showing the
cumbersome solution of (5.58), we note only that
3 ,
33 O(h/R, h/l).
= + O(h/R, h/l).
Thus,
we specify its potenConcerning the statically admissible electric field E
tial by
1
h2
= + (q A + q F )x q B (x2 ).
2
12
Then
E3 = (q A + q F ) + q B x = E3 ,
1
h2
E = F (q A + q F ), x + (q B ), (x2 )
2
12
= E + O(h/l).
constructed above satisfies
E)
Note that the statically admissible field (,
only the regular boundary conditions at the shell edge, exactly as in the
previous case.
Case ii) Construction of kinematically admissible field. The displacements
is calculated
w , w are given by (5.42) and (5.49), and the strain tensor
according to (5.30). As in the previous case it is easy to see that
= + O(h/R, h/l).

are given by (5.50)3 , and D


of D
3 is found from (5.57).
The components D
We pose at x = h/2 the following boundary condition
3 )|x=h/2 = 33 + e3 A .
(D
P
h

5.3. ERROR ESTIMATION AND COMPARISON

189

3 exists and D
3 = D3 + O(h/R, h/l).
Then D
Construction of statically admissible field. The electric potential is
given by
1
h2
0
x q B (x2 ),
=
h
2
4
= .
and E
It is easy to see that the conditions = 0 /2 are satisfied

is constructed in the same way as


and E = E + O(h/R, h/l). The tensor
= + O(h/R, h/l) can be proved
in case i), and the asymptotic formula
in a similar manner.
and
differ from those constructed by the 2-D theory
In both cases
by a quantity of the order h/R and h/l as compared with unity. We have
thus shown the asymptotic accuracy of the 2-D theory in the energetic norm
(5.51).
Special cases and comparison with the well-known theories
Transversely isotropic shell. This corresponds to the model of piezoceramic
shell with thickness polarization. As it was noted in Section 5.2, all the 2-D
tensors of electroelastic moduli of odd rank vanish, in particular e
= 0.
N
Since properties of such material are invariant with respect to the rotation
about the normal, we have
c
= cP1 a a + cP2 (a a + a a ),
P


c
= cN
+ cN
+ a a ),
1 a a
2 (a a
N

e3
= eP a ,
P

N = a .

The 2-D electric enthalpy becomes


2
N
0 = h {[cN

1 (A ) + 2c2 A A ]
2
h2 N 2

N 2
+ [c1 (B ) + cN
2 B B ] F },
12

(5.62)

in case i), and


1 = h {[cP1 (A )2 + 2cP2 A A ]

2
h2 N 2

P 0
+ [c1 (B ) + cN
},
2 B B ] + 2e A
12
h

(5.63)

in case ii). Note that the cross terms between the mechanical and electric
quantities are absent in (5.62).

190

CHAPTER 5. PIEZOELECTRIC SHELLS

We now express the coefficients cP1 , cP2 , eP through the 3-D electroelastic
moduli, whose components are denoted by means of the abbreviated indices
as described in Section 2.4. According to the formulae (5.32)
1
c12
cP2 = (c11
E ),
2 E
2
(c13
E)
12

cP1 = c11
(c11
E
E cE ),
c33
E
13 33
c
e
eP = e31 E 33 .
cE

(5.64)

Sometimes it is convenient to express cP1 , cP2 , eP through the other 3-D moduli,
am
abcd
for instance, smn
and dabc that enter the
E and d . These are related to sE
complementary energy density G(, E) in the following way: if among m and
abcd
n there are no indices 4,5,6, then smn
and dam = dabc ; if the indices
E = sE
abcd
and 12 dam = dabc ; if they are
4,5,6 are encountered once, then 12 smn
E = sE
1 mn
abcd
encountered twice, then 4 sE = sE [36]. For piezoceramics one can easily
show that
11 12 13 1
E
E
cE cE cE
s
s11 sE
13
12
E
E
,

s E
c13
c11
c12
s
s
=
13
11
12
E
E
E
33
13
13
E
E
E
cE cE cE
s13 s13 s33
and
31
E
E
e
s11 sE
d31
12 s13
E
E 31
d31 = sE
e
s
s
.
12
11
13
33
E
E
E
e
s13 s13 s33
d33

Using these relations and some algebra, we can transform (5.64) to


cP2 =

1
2sE
11 (1

+ )

cP1 =

,
2)

sE
11 (1

eP =

d31
E
s11 (1

E
where = sE
12 /s11 is Poissons ratio for the piezoceramic material. We now
N
calculate cN
1 and c2 . According to (5.32) we have

cN
2

1
= E
,
2s11 (1 + )

cN
1

+ 1
kp2
1
2
= E
,
s11 (1 2 ) 1 kp2

where kp2 is the piezoelectric coupling factor given by


kp2 =

2d231
.
(1 )T33 sE
11

5.3. ERROR ESTIMATION AND COMPARISON

191

In case i) (the piezoceramic shell with unelectroded face surfaces) the


constitutive equations read

N = h(cN
+ 2cN
1 A a
2 A ),
h3

M = (cN
+ 2cN
1 B a
2 B ).
12

(5.65)

In case ii) (the piezoceramic shell with electroded face surfaces) the constitutive equations take the form
N = h(cP1 A a + 2cP2 A + eP
M =

0
a ),
h

h3 N

(c B a + 2cN
2 B ).
12 1

(5.66)

It can be shown that in both cases the constitutive equations are asymptotically equivalent to the equations obtained in [10].1 In case i) the cross
terms between the mechanical and electric quantities in the electric enthalpy
of the classical theory vanish, and in order to describe the coupling piezoelectric effects we may need to construct a refined shell theory [50] and regard
it as the first-order approximation. In doing so the variational-asymptotic
method [6] again turns out to be much more effective as compared with other
asymptotic methods. However, we shall not go into the details of the refined
piezoelectric shell theories.
Piezoceramic shells with tangential polarization. We consider a piezoceramic
shell polarized along the tangents to the x2 co-ordinate lines of the middle
surface. We assume additionally that the co-ordinate lines x1 , x2 are simultaneously lines of principal curvatures of the middle surface. It is clear that
such a shell possesses the symmetry planes parallel to the middle surface,
and according to the note in Section 5.2
c
= c
,
N
P

e
= e
N
P ,

N = P .

Let us find out c


, e
and
P
P
P in terms of the 3-D electroelastic moduli. First of all we note the following useful relation
G( , 0, 0, Ea ) = max[  min W (ab , Ea )]


3 ,33

which follows from the definition of G( ab , Ea ) as the Legendre transformation of W (ab , Ea ) with respect to ab . Using this formula we can express
1

In [10] the measures of bending introduced by Goldenveizer and Novozhilov were used
instead of B or .

192

CHAPTER 5. PIEZOELECTRIC SHELLS

E
T
c
, e
and
P
P
P through s , d and
1
c
= (sE
) ,
P

ea
= c
da. ,
P
P

b
ab
a
ab
P = T d. eP ,

where the inverse of sE


should be understood in the following sense:

) . In accordance with the commonly established practice


cP s = (
for piezoceramic materials [20] the moduli sE
mn , dam and ab with the abbreviated indicial notation are always referred to the co-ordinate system, whose
z 3 -axis coincides with the axis of polarization. In terms of these moduli we
have
sE
sE
sE
1
33
13
11
12
22
=
c11
=

=
,
c
,
c
, c66
P
P
P
P = E ,

s44
E
E
E
E
d33 s11 d31 s13
d31 s33 d33 s13
e22
, e21
,
P =
P =

d15
11
16
22
22
21
e16
11
22
P = E ,
P = T d15 eP ,
P = T d33 eP d31 eP ,
s44
where
E
E 2
= sE
11 s33 (s13 ) ,

and the abbreviated indicial notation is used also for the 2-D moduli.
For the shell with unelectroded face surfaces the constitutive equations
(5.24) take the form
12
21
N11 = h(c11
P A11 + cP A22 eP F2 ),
22
22
N22 = h(c12
P A11 + cP A22 eP F2 ),
16
N12 = h(2c66
P A12 eP F1 ),
h3 11
M11 = (cP B11 + c12
P B22 ),
12
h3
B11 + c22
M22 = (c12
P B22 ),
12 P
h3
M12 = 2c66
B12 ,
12 P
G1 = h(2e16 A12 + 11
P F1 ),
21
22
G2 = h(eP A11 + eP A22 + 22
P F2 ).

For the shell with electroded face surfaces we obtain from (5.26)
12
22
N11 = h(c11
N22 = h(c12
P A11 + cP A22 ),
P A11 + cP A22 ),
h3 12
h3 11
12
M11 = (cP B11 + cP B22 ), M22 = (cP B11 + c22
P B22 ),
12
12

5.4. FREQUENCY SPECTRA OF CIRCULAR PLATES


N12 =

2hc66
P A12 ,

M12

193

h3 66
= 2cP B12 .
12

In both cases the constitutive equations are asymptotically equivalent to


the equations obtained in [50].2 In the second case the cross terms between
the mechanical and electric quantities in the electric enthalpy of the classical
theory vanish, so that the refined shell theory [50] might become essential.
Problems
1. Find the solution of the equation (5.58) and show that
3 ,
33
O(h/R, h/l).
2. Construct the kinematically and statically admissible fields for piezoelectric shells with the electroded face surfaces.
3. For anisotropic piezoelectric crystals show that
a
b
1
ab
a
c
= (sE
= c
da. , ab
) , eP
P = T d. eP .
P
P

5.4

Frequency spectra of circular plates

In this section we illustrate the application of the theory to the problem


of axisymmetric longitudinal vibrations of piezoceramic circular plates with
thickness polarization. We analyze the resonant and antiresonant vibrations
of plates with the three methods of electrode arrangements enumerated at
the beginning of the chapter.
Unelectroded face surfaces. According to (5.62) the coupling effect disappears in the electric enthalpy so that the problem reduces to the purely
mechanical one. Introducing the polar co-ordinates %, , we calculate the
measures of extension for axisymmetric longitudinal vibrations
A%% = u%,% ,

A% = 0,

A =

u%
,
%

where u% (%, t) is the only nonvanishing component of the displacement. Since


the latter is independent of , the mechanical part of the action functional,
up to the unimportant factor h, reads
Z t1 Z r
u2
u% 2
N %
N 2
[u 2% cN
(u
+

2c
)

2c
u
%,%
1
2 %,%
2 2 ]%d% dt,
%
%
t0
0
2

In [50] the Goldenveizer-Novozhilov measures of bending were used.

194

CHAPTER 5. PIEZOELECTRIC SHELLS

where r is the radius of the plate. The Euler equation of this functional takes
the form
N

u% = (cN
1 + 2c2 )(u%,%% +

u%,% u%
2 ),
%
%

(5.67)

combined with the traction-free boundary condition


N
N
(cN
1 + 2c2 )u%,% + c1

u%
= 0 at % = r.
r

(5.68)

We seek solutions of (5.67) in the form


u% = ueit .

(5.69)

Taking into account that


cN
1

2cN
2

1 1
kp2
1
2
= E
,
s11 (1 2 ) 1 kp2

cN
1

+ 1
kp2
1
2
= E
,
s11 (1 2 ) 1 kp2

and introducing the dimensionless quantities


q
%
2
y = , = r sE
11 (1 ),
r

(5.70)

we transform (5.67) and (5.68) to


1
2
1
u00 + u0 + ( 2 2 )
u = 0,
y

u0 + 2 u = 0 at y = 1,

(5.71)

where the prime denotes the derivative with respect to y and


2 =

1 1
kp2
2
,
1 kp2

+ 1
kp2
2
.
1 kp2

(5.72)

The non-singular solution of (5.71) is given by

u = aJ1 ( y),

where should be found from the equation

J0 ( ) = (1 2 )J1 ( ).

(5.73)

5.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

195

Taking = 0.35, kp = 0.6,3 we find the first three roots of (5.73) to be equal
to 2.563, 6.381, 10.1, respectively.
Fully electroded face surfaces. In this case the action functional reduces
to
Z t1 Z r
u2%
u%
[u 2% cP1 (u%,% + )2 2cP2 u2%,% 2cP2 2
%
%
0
t0
u%
0
2eP (u%,% + )]%d% dt,
h
%
which leads to the following Euler equation

u% = (cP1 + 2cP2 )(u%,%% +

u%,% u%
2 ),
%
%

(5.74)

and the traction-free boundary condition


(cP1 + 2cP2 )u%,% + cP1

u%
0
+ eP
= 0 at % = r.
r
h

(5.75)

Letting the voltage 0 (t) depend harmonically on t


0 (t) = 0 eit ,
we seek solutions of (5.74) and (5.75) in the form (5.69). Since
cP1 + 2cP2 =

1
,
2)

sE
11 (1

cP1 =

,
2)

sE
11 (1

eP =

d31
E
s11 (1

equations (5.74), (5.75), in terms of the dimensionless variable y, can be


transformed to
1
1
u00 + u0 + (2 2 )
u = 0,
y
y
0
u0 + u + d31 (1 + ) r = 0 at y = 1,
h
The non-singular solution is found to be
u = aJ1 (y),

(5.76)

where a is equal to
a=
3

d31 (1 + )r
0
.
(1 )J1 () + J0 () h

These are the material constants of the piezoceramic PZT-5 [9].

(5.77)

196

CHAPTER 5. PIEZOELECTRIC SHELLS

Thus, we find the resonant frequencies from the equation


(1 )J1 () = J0 (),

(5.78)

which, for = 0.35, gives the first three roots as 2.08, 5.399, 8.578.
To find the antiresonant frequencies we should determine the surface
charges on electrodes or, alternatively, the current. We use the equation
(5.50) for D3 , which, in our case, reduces to
u%
0
+ eP (u%,% + ).
D3 = 33
P
h
%
Taking into account that 0 and u% depend on t harmonically, we eliminate
their common factor eit to obtain the amplitude of D3
0
u
3 = 33
D
+ eP (
u,% + ).
P
h
%
2
T
from (5.76) gives
It is easy to check that 33
P = 33 (1 kp ). Substitution of u

d31

3 = T33 (1 kp2 ) 0 +
D
a J0 ( %).
E
h
r
s11 (1 ) r
Now we calculate the amplitude of the surface charge on one electrode by
3 over S
integrating D
Z
kp2
0
3
2 T
2

aJ1 ()].
D da = r 33 [(1 kp )
h
rd31
S
Remembering (5.77), we find that the antiresonant frequencies, for which the
surface charges should vanish, are determined as roots of the equation
J0 () = [1 (1 + )

kp2
]J1 ().
1 kp2

(5.79)

For = 0.35, kp = 0.6, we find the first three roots of this equation to be
2.449, 5.54, 8.666.
Partially electroded face surfaces. Suppose that the electrodes cover
only the rings r0 % r of the face surfaces. On the electrodes the values of
the electric potential 0 /2 are specified. For the axisymmetric longitudinal
vibrations the action functional, up to an unimportant factor, is the sum of
the following integrals
Z t1 Z r
Z t1 Z r0
u2
u% 2
N %
2
N 2
u % %d% dt
[cN
(u
+
)
+
2c
u
+
2c
%,%
2 2 ]%d% dt
1
2 %,%
%
%
t0
0
t0
0
Z t1 Z r
2
u%
0
u%
u%

[cP1 (u%,% + )2 + 2cP2 u2%,% + 2cP2 2 + 2eP (u%,% + )]%d% dt.


%
%
h
%
t0
r0

5.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

197

Varying this functional, we obtain the Euler equation


u%,% u%
2 ) for % (0, r0 ),
%
%
u%,% u%
P
P

u% = (c1 + 2c2 )(u%,%% +


2 ) for % (r0 , r),
%
%

u% = (cN
1 + 2c2 )(u%,%% +

(5.80)

the jump conditions at % = r0


u% |r0 = u% |r0 + ,
N
N
[(cN
1 + 2c2 )u%,% + c1 u% /%]|r0

(5.81)

0
= [(cP1 + 2cP2 )u%,% + cP1 u% /% + eP ]|r0 + ,
h
and the traction-free boundary condition at % = r
[(cP1 + 2cP2 )u%,% + cP1 u% /% + eP

0
]|r = 0.
h

(5.82)

The voltage 0 (t) is assumed to depend harmonically on t, 0 (t) = 0 eit ,


so that solutions of (5.80)-(5.82) can be sought in the form (5.69). Again, in
terms of the dimensionless variable y introduced in (5.70) this system can be
transformed to the differential equations
2
1
u00 + u0 + ( 2
y

1
u00 + u0 + (2
y

1
)
u = 0 for y (0, r),
y2
1
)
u = 0 for y (
r, 1),
y2

where r = r0 /r, the jump conditions at y = r


u|r = u|r+ ,
[2 u0 +
u/y]|r = [
u0 + u/y + d31 (1 + )

0
r]|r+ ,
h

and the traction-free boundary condition at y = 1


[
u0 + u/y + d31 (1 + )

0
r]|y=1 = 0.
h

For the part of the plate without electrodes the solution is given by

u = a1 J1 ( y),

while for the part covered by the electrodes we have


u = a2 J1 (y) + a3 Y1 (y).

198

CHAPTER 5. PIEZOELECTRIC SHELLS

The constants a1 , a2 , a3 can be determined from the jump conditions at y = r0


and the boundary condition at y = 1. Substituting the formulae for u into
them, we get the system of linear equations
3
X

Cij aj = bi ,

i = 1, 2, 3,

j=1

where Cij and bi are given by


r
C11 = J1 ( ),

C12 = J1 (
r),

C13 = Y1 (
r),

r
r
r
C21 = (2 )J1 ( ) + 2 J0 ( ),

C22 = (1 )J1 (
r)
rJ0 (
r),
C23 = (1 )Y1 (
r)
rY0 (
r),
C31 = 0, C32 = (1 )J1 () J0 (),
C33 = (1 )Y1 () Y0 (),
0
0
b1 = 0, b2 = d31 (1 + )r r, b3 = d31 (1 + )r .
h
h
After finding ai we can determine the amplitude of D3 by the formula
u
d31
3 = T (1 k 2 ) 0 +
(
u,% + )
D
33
p
E
h
%
s11 (1 )
d31
0
[a2 J0 (%) + a3 Y0 (%)],
= T33 (1 kp2 ) + E
h
s11 (1 )
where = /r. Then the amplitude of the total charge on one of the electrodes is equal to
Z
3 da = r2 T 0 {(1 r2 )(1 k 2 ) k 2 [
D
J1 (
r))
33
p
p a2 (J1 () r
h
S1
+a
3 (Y1 () rY1 (
r))]},
where a
i is the solution of the system
3
X

Cij a
j = bi ,

i = 1, 2, 3,

j=1

with
b1 = 0,

b2 = (1 + )
r,

b3 = 1 + .

5.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

199

Figure 5.2: Resonant and antiresonant frequencies of the partially electroded


circular plate.
According to the solution of this problem the resonant frequencies are the
roots of the determinantal equation
det Cij = 0.

(5.83)

The antiresonant frequencies should be found from the condition that the
total charge vanishes giving
(1 r2 )

1 kp2
=a
2 [J1 () rJ1 (
r)]
kp2

(5.84)

+
a3 [Y1 () rY1 (
r)].
For = 0.35, kp = 0.6 the first three roots of equations (5.83) and (5.84) as
functions of the parameter r = r0 /r are numerically evaluated and the results
are plotted in Figure 5.2, where the solid lines correspond to the resonant
frequencies and the dashed lines to the antiresonant ones. One can see that
for r = 1 and r = 0 these curves give the values found previously.
Problems
1. Plot the first three resonant frequencies of (5.78) as functions of .
2. Plot the first three antiresonant frequencies of (5.79) as functions of
taking kp = 0.6.

200

CHAPTER 5. PIEZOELECTRIC SHELLS

3. The electromechanical coupling coefficient is defined by


kd2 =

a2 r2
,
a2

where r is the resonant frequency and a the antiresonant frequency.


Plot this coefficient as a function of r for partially electroded plates.

5.5

Frequency spectra of cylindrical shells

In this section we apply the piezoelectric shell theory to the problem of axisymmetric vibrations of piezoceramic circular cylindrical shells with thickness
polarization. We analyze the resonant and antiresonant vibrations of shells
with different electrode arrangements.
Unelectroded face surfaces. Consider a finite circular cylindrical shell
of the thickness h, the radius R, and the length 2L, referred to the same
co-ordinates as in Section 3.4. The shell is assumed to be made of a piezoceramic material with thickness polarization. We study the axisymmetric
vibrations, for which the component u2 = 0 and the components u1 , u of the
displacements do not depend on x2 . Then the measures of extension and
bending are given by
u
A11 = u1,1 , A12 = A21 = 0, A22 = ,
R
(5.85)
B11 = u,11 , B12 = B21 = 0, B22 = 0.
Substituting these formulae into (5.62), we obtain the expression of the elec 0 in terms of displacements. Note that the last term N F2
tric enthalpy

of 0 can be neglected due to the absence of the piezoelectric coupling effect


in this case. Since u1 , u do not depend on x2 , the 2-D action functional takes
the form
Z t1 Z L
u 2
2
{(u 21 + u 2 ) [cN
) + 2cN
J =Rh
1 (u1,1 +
2 u1,1
R
t0
L
2
h2 N
N u
2
1
+ 2c2 2 + (c1 + 2cN
(5.86)
2 )u,11 ]} dx dt.
R
12
Thus, we obtain the purely mechanical problem. The Euler equations of this
functional read
u,1

u1 = cN
) + 2cN
(5.87)
1 (u1,11 +
2 u1,11 ,
R
cN
u
h2 N
1
N u

u = (u1,1 + ) c2 2 (c1 + 2cN


2 )u,1111 .
R
R
R
12

5.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

201

The boundary conditions corresponding to the clamped, fixed or free edges


can be derived. We restrict ourselves to the free edge boundary conditions,
which are the natural boundary conditions of the functional (5.86)
u
) + 2cN
cN
2 u1,1 = 0,
1 (u1,1 +
R
u,11 = u,111 = 0 at x1 = L.
For free vibrations, we seek the solution of (5.87) in the form
u1 = u1 eit ,

u = ueit ,

(5.88)

where is an eigenfrequency. In terms of the dimensionless quantities


q
x1
1
2 ),
(1

=
,
(5.89)
= R sE
11
R
the equations of motion reduce to
2 u001 +
u0 + 2 u1 = 0,

u01 + 2 u + ? u0000 2 u = 0,

(5.90)

where the parameters 2 and have been introduced in (5.72) and


kp2
1 h2 2
1 h2 1 1
2
? =
=
.
12 R2
12 R2 1 kp2

(5.91)

The prime is used to denote the derivative with respect to 1 . In a similar


manner as in Section 3.6 one can show that the symmetric solution of (5.90)
reads
u1 =

3
X

ai sin i ,

i=1

u =

3
X

ai

i=1

i
cos i 1 ,
i

(5.92)

where ai are unknown constants, 2i are the roots of the equation


(2 2 + 2 )(? 4 + 2 2 ) + 2 2 = 0,

(5.93)

i = 2 2i + 2 .

(5.94)

and

For each there are three different roots of equation (5.93), whose behaviour
is qualitatively similar to that shown in Figure 3.18. However, here the cutoff frequency is equal to . Substituting (5.92) into the free edge boundary
conditions at 1 = l = L/R
2 u01 +
u = 0,

u00 = u000 = 0,

(5.95)

202

CHAPTER 5. PIEZOELECTRIC SHELLS

we get the following system of linear homogeneous equations


3
X

Cij aj = 0,

i = 1, 2, 3,

j=1

with the following components of the matrix Cij


2
cos j l,
j
C2j = j j cos j l,
C3j = j 2j sin j l.
C1j =

(5.96)

The determinantal equation


det Cij = 0
yields the eigenfrequencies of free vibrations. This equation is numerically
evaluated for = 0.35, kp = 0.6, h/R = 0.1 and the first two frequencies are
plotted in Figure 5.3 as functions of the ratio l = L/R.

Figure 5.3: Eigenfrequencies versus l of piezoceramic unelectroded cylindrical shells with free edges ( = 0.35, kp = 0.6, h/R = 0.1).
The fundamental frequency as well as the overtones of piezoceramic shells
with other boundary conditions can be evaluated in exactly the same manner
as in Section 3.6.

5.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

203

Fully electroded face surfaces. Taking (5.85) into account, we write down
the 2-D action functional in the form
Z t1 Z L
u
{(u 21 + u 2 ) [cP1 (u1,1 + )2 + 2cP2 u21,1
J =Rh
R
L
t0
u2
u 0 (t) h2 N
2
1
+ 2cP2 2 + 2eP (u1,1 + )
+ (c1 + 2cN
2 )u,11 ]} dx dt.
R
R h
12
Consequently, the Euler equations read
u,1
) + 2cP2 u1,11 ,
(5.97)
R
cP
u
u
0 (t) h2 N

u = 1 (u1,1 + ) cP2 2 eP
(c1 + 2cN
2 )u,1111 .
R
R
R
hR
12

u1 = cP1 (u1,11 +

For the free edges we have the following natural boundary conditions
u
0 (t)
) + 2cP2 u1,1 + eP
= 0,
R
h
u,11 = u,111 = 0 at x1 = L.

cP1 (u1,1 +

(5.98)

Let the voltage 0 (t) be a harmonic function of t: 0 (t) = 0 eit . We seek


solutions of (5.97) and (5.98) in the form (5.88). Equations (5.97) and (5.98),
in terms of the dimensionless variables (5.89), reduce to
u001 + u0 + 2 u1 = 0,
0
u01 + u + ? u0000 + R 2 u = 0,
h

(5.99)

0
= 0,
h
u00 = u000 = 0 at 1 = l,

(5.100)

and
u01 + u + R

where = d31 (1 + ).
It is easy to see that the symmetric solutions of (5.99) and (5.100) are
given by
u1 =

3
X
i=1

ai sin i ,

X i
R0
u =
+
ai
cos i 1 ,
h(1 2 ) i=1 i

where 2i are the three roots of the cubic equation


(2 + 2 )(? 4 + 1 2 ) + 2 2 = 0,

(5.101)

204

CHAPTER 5. PIEZOELECTRIC SHELLS

i are expressed through 2i in the following way


i = 2i + 2 ,
and ai should be found as the solution of the system of linear equations
3
X

Cij aj = bi ,

i = 1, 2, 3.

j=1

The elements of the matrix Cij turn out to be identical in form with (5.96),
while bi are given by
b1 = R

0 1 + 2
,
h 1 2

b2 = b3 = 0.

In order to calculate the amplitude of the total charge on one of the


electrodes, we should know the amplitude of the electric induction D3 which,
in our case, is equal to
d31
u
0
3 = T33 (1 kp2 ) 0 +
(
u1,1 + ) = T33 (1 kp2 )
D
E
h
R
h
s11 (1 )
3
X
d31
d31
0
i
+
) cos i 1 .
E
a
(
+
i
i
E
i
s11 (1 ) h(1 2 ) s11 (1 )R i=1
Integration of this formula gives the amplitude of the total charge on one of
the electrodes
Z
k 2 (1 + )l
3 da = 2R2 T 0 [(1 k 2 )2l p
D
33
p
h
1 2
S
3
X
i sin i l
2
)
],
+ kp (1 + )
a
i (i +
i i
i=1
where a
i are the solution of the following system
3
X

Cij a
j = bi ,

i = 1, 2, 3,

j=1
2
b1 = 1 + ,
1 2

b2 = b3 = 0.

Thus, the resonant frequencies correspond to the roots of the equation


det Cij = 0,

(5.102)

5.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

205

while the antiresonant frequencies should be found from the condition of


vanishing charges on electrodes giving
(1

kp2 )2l

3
X
kp2 (1 + )l
i sin i l
2
)
.
+
= kp (1 + )
a
i (i +
2
1

i
i
i=1

(5.103)

For = 0.35, kp = 0.6, h/R = 0.1 the first two roots of equations (5.102) and
(5.103) as functions of the parameter l = L/R are numerically evaluated and
the results are plotted in Figure 5.4, where the solid lines correspond to the
resonant frequencies and the dashed lines to the antiresonant ones.

Figure 5.4: Resonant and antiresonant frequencies of piezoceramic fully electroded cylindrical shells with free edges ( = 0.35, kp = 0.6, h/R = 0.1).
Partially electroded face surfaces. Assume that the portions |x| < L0
of the shell face surfaces are covered by electrodes and the values 0 /2 of
the electric potential are specified on them. For axisymmetric harmonic vibrations the 2-D governing equations, in terms of the dimensionless variables
(5.89), are of the form (5.99) in the electroded region | 1 | < l0 = L0 /R, and
of the form (5.90) in the remaining region l0 < | 1 | < l. The general symmetric solution of (5.99) is given by (5.101), while that in the unelectroded
region should be modified as follows
u1 =

6
X

aj sin j +

j=4
6
X

9
X

ak cos k 1 ,

k=7
9
X

k
j
cos j 1
sin k 1 ,
u =
aj
ak

j
k
j=4
k=7

(5.104)

206

CHAPTER 5. PIEZOELECTRIC SHELLS

for l0 < < l, and determined for l < < l0 so that u1 , u are odd
and even functions, respectively. In (5.104) 2j , j = 4, 5, 6 are the three roots
of the cubic equation (5.93) and 7 = 4 , 8 = 5 , 9 = 6 . The j are
determined through 2j according to (5.94).
The jump conditions at 1 = l0 read
R0
)|l0 = (2 u01 +
u)|l0 + ,
h
u00 |l0 = u00 |l0 + , u000 |l0 = u000 |l0 +
= u1 |l0 + , u|l0 = u|l0 + , u0 |l0 = u0 |l0 + .
(
u01 + u +

u1 |l0

Finally, we have to satisfy the free edge boundary condition (5.95). Substituting the general solution containing nine unknown constants ai into these
nine conditions, we obtain the system of nine linear equations
9
X

Cij aj = bi ,

i = 1, . . . , 9.

(5.105)

j=1

The elements of the 9 9 matrix Cij are given by


2
2
2
cos i l0 , C1j = cos j l0 , C1k =
sin k l0 ,
i
j
k
i i
j j
k k
C2i =
cos i l0 , C2j =
cos j l0 , C2k =
sin k l0 ,

j 2j
i 2i
k 2k
C3i =
sin i l0 , C3j =
sin j l0 , C3k =
cos k l0 ,

C4i = sin i l0 , C4j = sin j l0 , C4k = cos k l0 ,


i
j
k
C5i =
cos i l0 , C5j =
cos j l0 , C5k =
sin k l0 ,
i
j
k
i
j
k
C6i = sin i l0 , C6j =
sin j l0 , C6k =
cos k l0 ,

2
2
C7i = 0, C7j =
cos j l, C7k = sin k l,
j
k
C8i = 0, C8j = j j cos j l, C8k = k k sin k l,
C9i = 0, C9j = j 2j sin j l, C9k = k 2k cos k l.
C1i =

In these formulae the index i runs from 1 to 3, j from 4 to 6, and k from 7


to 9. The bi in the right-hand side of (5.105) have only two elements which
are not zero, namely
b1 =

R0 1 + 2
,
h
1 2

b5 =

R0
.
h(1 2 )

5.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

207

Figure 5.5: First resonant and antiresonant frequency of piezoceramic partially electroded cylindrical shells versus l0 ( = 0.35, kp = 0.6, h/R = 0.1, l =
1).
3,
After finding ai we calculate the amplitude of the electric induction D
and then the total charge on the electrodes. One can show that the latter is
given by
Z
k 2 (1 + )l0
3 da =2R2 T33 0 [(1 kp2 )2l0 p
D
h
1 2
S1
3
X
i sin i l0
2
)
],
(5.106)
+ kp (1 + )
a
i (i +
i
i
i=1
where a
i are the solution of the following system
9
X

Cij a
j = bi ,

i = 1, . . . , 9,

j=1
2

b1 = 1 + ,
1 2

b5 =

1
,
1 2

bi = 0, i 6= 1, 5.

Thus, the resonant frequencies are determined as roots of the equation


det Cij = 0,

(5.107)

while the antiresonant frequencies should be found from the condition that
the right-hand side of (5.106) vanishes. For = 0.35, kp = 0.6, h/R =
0.1, l = 1 the first resonant and antiresonant frequencies versus the length
of the electrodes l0 = L0 /R are plotted in Figure 5.5, where the solid lines
correspond to the resonant frequency and the dashed lines to the antiresonant
one.

208

CHAPTER 5. PIEZOELECTRIC SHELLS


Problems

1. Study the frequency spectra of the piezoceramic cylindrical shell with


thickness polarization and with fixed edges.
2. Do the same with the clamped edges.
3. Calculate the electromechanical coupling coefficient defined by
kd2 =

a2 r2
,
a2

and plot it as the function of l for fully electroded cylindrical shells.

Chapter 6
Piezoelectric rods
6.1

One-dimensional equations

Geometry of a piezoelectric rod and electrode arrangement. Consider a domain B in the three-dimensional Euclidean space E obtained by
moving a plane connected figure S along a smooth curve c(x) so that c(x)
always remains orthogonal to S and cuts it in a fixed point (see Figure 4.1).
A linear piezoelectric body occupying the domain B in its stress-free undeformed state is called a piezoelectric rod, the curve c(x) its central line, and
S its cross section.
Let z i = ri (x) be the equation of the central line, with x being the arclength. We denote by t the unit tangent vector to c(x), and by t1 , t2 the unit
vectors orthogonal to each other and to t so that they are rigidly mounted
with the cross section and move together with S as it moves along the central
line. We assume that t1 , t2 and t form a positive-oriented triad. The domain
B E is specified by the equation of the form
z i (xa ) = ri (x) + ti (x)x ,

(6.1)

where ti (x) are cartesian components of the vectors t (x). The co-ordinates
x take values in a connected domain S R2 . We assume that the point

with
R co-ordinates x = 0 coincides with the centroid of the domain S, so
x da = 0.
S
Let S denote the boundary of the cross section. We consider the three
methods of electrode arrangement encountered most often:
i) There are no electrodes on the lateral boundary of the rod (fully unelectroded lateral surface). The edges of the rod are coated by two
electrodes and the values of the electric potential are prescribed on
209

210

CHAPTER 6. PIEZOELECTRIC RODS


them
= 0 (t)/2,

for x = 0, L.

(6.2)

ii) The lateral surface S (0, L) is partially coated by electrodes in form


of strips. Thus, we assume that S is decomposed into open curves
(1)
(n)
ce , . . . , ce (where there are electrodes) and the remaining part cd .
(i)
For xa ce [0, L] on the electroded portions of the lateral boundary
values of the electric potential are prescribed
= (i) (t),

i = 1, . . . , n.

(6.3)

On the unelectroded portion of the lateral boundary as well as on the


edges of the rod the surface electric charge should vanish.
iii) Along the central line of the rod the lateral surface is decomposed
successively into partially electroded and fully unelectroded portions.
This case can be regarded as the mixed situation of the two cases
considered above.
Kinematics of a one-dimensional piezoelectric rod. Similar to the
2-D piezoelectric shell theory, the number of unknown functions specifying
the kinematics of piezoelectric rods in the one-dimensional setting depends
on the location of electrodes. In case i) the kinematics of a piezoelectric
rod is completely specified by the five unknown functions, namely the three
components of the displacements u(x, t) of the central line, the function
(x, t) describing the small rotation of the cross section1 and (x, t) being
the 1-D electric potential. These functions are assumed to be continuous
and as many times differentiable as required. The measures of elongation,
bending and twist are defined by (cf. (4.3),(4.8) and (4.9))
= u0 + u ,
0
.
= e.
. $ + e. ,

(6.4)

= 0 + ,
where
0
.
= e.
. u $u + e. u.

(6.5)

In these formulae and $ correspond to the curvatures and torsion of


the rod in its natural state, respectively, u = t u and u = t u are the
1

We use for the rotation angle instead of , since the latter is occupied by the electric
potential.

6.1. ONE-DIMENSIONAL EQUATIONS

211

components of the displacements referred to the basis {t , t}. As in Chapter


4, we use e to denote the two-dimensional permutation symbols (e11 =
e22 = 0, e12 = e21 = 1). The one-dimensional electric field F is obtained by
differentiating the electric potential
F = 0 .

(6.6)

In case ii) the number of kinematically independent functions for a piezoelectric rod is reduced to four the displacements u(x, t) and the rotation
(x, t), as for the elastic rods. The one-dimensional electric enthalpy of the
rod depends on the measures of elongation, bending and twist, referred to
as state variables in this case. Finally, in case iii) we must combine the two
types of kinematics described above. In the fully unelectroded portion of the
lateral surface we need the kinematics of the type i), while in the partially
electroded portion of the lateral surface that of the type ii).
Since u and depend on t we introduce the following quantities
v = u (velocity),

(acceleration),
a=u

,

to measure their time rates. The latter quantities describe the angular velocity and acceleration of the torsional motion of the rod. The time rates of
do not play any role in the one-dimensional theory.
Variational principles and boundary-value problems
Case i) For a piezoelectric rod with unelectroded lateral surface the onedimensional action functional reads
Z t1 Z L
J[u, , ] =
( ) dx dt,
(6.7)
t0

where is the kinetic energy density and the electric enthalpy. The kinetic
energy density is a quadratic form of u and
1
1
= (u u + u 2 ) + 2 ,
2
2

(6.8)

where is the mass density per unit length, and is a constant to be determined later. The electric enthalpy is a quadratic form of the measures of
elongation, bending, and twist and of the electric field, which, in the general
case, is given by
0 (, , , F ).
=

212

CHAPTER 6. PIEZOELECTRIC RODS

Specifying u and at t = t0 and t = t1


u|t=t0 = u0 ,
|t=t0 = 0 ,

u|t=t1 = u1 ,
|t=t1 = 1 ,

we formulate Hamiltons variational principle for piezoelectric rods as follows:


, rotation and electric potential correspond to
the true displacements u
the stationary points of the action functional (6.7)
J = 0.
To derive the 1-D equations of motion from this variational principle let
us calculate the variation of the functional (6.7)
Z t1 Z L
T
J =
[
(u u + u
u)
+
t0

M M G 0 ] dx dt.

(6.9)

Here the following notations are introduced

, M =
,

M=
, G=
.

T =

(6.10)

We call T the tension, M the bending moments, M the twisting moment


and G the 1-D electric induction. If the electric enthalpy is known, equations
(6.10) can be regarded as the constitutive equations for piezoelectric rods.
According to (6.4) the variations of , and are given by
= u0 + u ,
0
0
.
0
= u00 + e.
. ($u ) + ( u) + e. $u
.
+ $2 u e.
. $u + e. ,

= 0 e u0 $u .
Substituting these formulae into (6.9) and assuming the regularity of all the
quantities under the integral sign, we integrate the latter by parts for the
variations vanishing at x = 0, L to obtain
Z t1 Z L
J =
{(
u + T 0 + M0 e M $ )u + [
u T
+

t0

M00

0
0
0

e.
. (M $ + (M $) (M ) ) (M $ M )$]u
+ (
+ M 0 e M ) + G0 } dx dt = 0.

6.1. ONE-DIMENSIONAL EQUATIONS

213

Since the variations of u, u , and are arbitrary inside the region (0, L)
(t0 , t1 ), we conclude that
u = T 0 + M0 e M $ ,
0
0
0
u = T + M00 e.
. [M $ + (M $) (M ) ]

(M $ M )$,

(6.11)

= M 0 e M ,
G0 = 0.
These are the 1-D equations of motion of the piezoelectric rod without electrodes on the lateral boundary. Substituting the constitutive equations (6.10)
into (6.11), we obtain five differential equations with respect to the five unknown functions u, u , and .
Letting the variations of the unknown functions be arbitrary at x = 0, L,
we derive, in a similar manner as in Chapter 4, the free edge boundary
conditions:
T = 0, M = 0, M0 = 0, M = 0.
(6.12)
Besides, the electric potential satisfies the conditions
= 0 (t)/2,

for x = 0, L.

It is not difficult to write down the boundary conditions for the clamped or
fixed edge.
Case ii) For a piezoelectric rod with partially electroded lateral surface the
one-dimensional action functional depends only on the displacements u and
rotation
Z t1 Z L
J[u, ] =
( ) dx dt,
(6.13)
t0

1 (, , ) the
where is the kinetic energy density given by (6.8) and =
electric enthalpy, which depends only on the measures of elongation, bending
and rotation correspond to the staand twist. The true displacements u
tionary points of the functional (6.13). Thus, this variational principle leads
to the same equations of motion and boundary conditions as in the theory
of elastic rods (cf. equations (4.16) and (4.17)). However, the constitutive
equations will be changed.
Case iii) Let c0 be a portion of the central line, the lateral surface of which is
fully unelectroded, and c1 the remaining part. Assume, for definiteness, that

214

CHAPTER 6. PIEZOELECTRIC RODS

c0 lies strictly inside c(x) and c0 is the boundary between these two regions.
The 1-D action functional can be presented as a sum of two integrals
Z t1 Z
0 (, , , F )] dx dt
[
J[u, , ] =
t0
c0
(6.14)
Z t1 Z

[ 1 (, , )] dx dt.
+
t0

c1

We admit the displacement fields with discontinuous first and second derivatives at the boundary c0 . However we require that u , u and as well as
the first derivative u0 should be continuous there. Concerning the 1-D admissible electric potentials we require that they vanish at the boundary c0 .
The true diplacement field, rotation and electric potential correspond to the
stationary points of the functional (6.14). Varying this action functional
with the vanishing variations at the boundary c0 , we obtain the equations
of motion (6.11) in the region c0 and (4.16) in c1 . The boundary conditions
at the free edge of the shell x = 0, L are given by (6.12). For the variations
not vanishing at the boundary c0 we obtain the following jump conditions
[[T ]] = 0,

[[M ]] = 0,

[[M0 ]] = 0,

[[M ]] = 0,

where [[.]] denotes the jump of the corresponding quantity across the point
x0 c0 : [[A]] = A+ A , the indices + and indicating the limiting values
of A as x x0 0. Since the electric potential vanishes at c0 there is no
jump condition for the electric induction field.
1-D electric enthalpy.
Case i) We assume that the rod is made of a linear homogeneous piezoelectric
material. It turns out that, within the first-order approximation, the electric
enthalpy density of the rod is given by
1
0 = (cK |S| 2 + A + 2B + C2
2
(6.15)
2
2eK |S|F K |S|F ).
The coefficient cK |S| plays the role of the effective Youngs modulus, A
correspond to the flexural rigidities, C the torsional rigidity, K |S| the
1-D effective dielectric permittivity. The remaining coefficients B and eK |S|
characterize the coupling effects between bending and torsion and between
elongation and 1-D electric field, respectively.
Case ii) For homogeneous piezoelectric rods the 1-D electric enthalpy is given
by
1
1 = (cK |S| 2 + cK I ) + (, , , (i) ),
(6.16)
2

6.2. ASYMPTOTIC ANALYSIS

215

where (, , , (i) ) is a quadratic form with respect to , and and


can be determined only after solving the so-called cross section problem.
The latter can only be analyzed in particular cases, some of which will be
presented in Section 6.3.
Problems
1. Derive the balance equation of energy for piezoelectric rods. Using
this equation, prove the uniqueness of solutions of the boundary-value
problems.
2. Formulate the variational principles for the eigenvalue problems of
determining the resonant and antiresonant vibrations of piezoelectric
rods.
3. Prove the orthogonality of the eigenfunctions of the resonant or antiresonant vibrations of piezoelectric rods.
4. Show that the equations of motion for a naturally straight and untwisted piezoceramic rod with longitudinal polarization and without
electrodes on the lateral surface break into the equations of longitudinal, flexural and torsional vibrations. Find the corresponding equations
in terms of u, u and .

6.2

Asymptotic analysis

Three-dimensional action functional. Assume that the domain B defined by (6.1) is occupied by a linear piezoelectric rod in its stress-free undeformed state. The 3-D co- and contravariant components of the metric
tensor in the chosen curvilinear co-ordinate system x , x can be calculated
with the help of the formulae (4.20)-(4.23). Denote by L the total length
of the central line c(x), by h the diameter of S, and by R the characteristic
radius of curvatures and torsion of the rod. When
h/L  1 and h/R  1,

(6.17)

the rod is said to be thin. Concerning the location of electrodes we restrict


ourselves to the three cases i)-iii) described in the previous section. Since case
iii) is simply the combination of the first two cases, the asymptotic analysis
will be done for the first two cases only.

216

CHAPTER 6. PIEZOELECTRIC RODS

In accordance with the variational principle (2.45), (2.47) the true displacement field wi and the electric potential are sought as stationary points
of the action functional
Z t1 Z L Z

[T W (ab , Ea )] g da dx dt
I=
(6.18)
t0

under the constraints (6.2) in case i), and the constraints (6.3) in case ii),
where T and W are the 3-D kinetic energy and electric enthalpy densities,
respectively. The kinetic energy density T is given by (4.25), while the electric
enthalpy W (ab , Ea ) is the quadratic form (5.28).
The problem is to replace the three-dimensional action functional (6.18)
by an approximate one-dimensional functional for a thin rod, whose kinematics is specified by functions depending only on the longitudinal co-ordinate x
and time t. The possibility of passing from the three- to the one-dimensional
problem is related to the smallness of h/R and h/l, where l is the characteristic scale of change of the electroelastic state in the longitudinal direction
x. We also restrict ourselves to those dynamical processes for which the
inequality
h
1
(6.19)
c
holds, with the characteristic scale of change of the function wi in time and
c the minimal velocity of plane waves in the piezoelectric material. According
to (6.19) only low-frequency vibrations of the piezoelectric rod are considered
here. By using the variational-asymptotic method, one-dimensional action
functionals will be constructed in which terms of the order h/R and h/l are
neglected compared with unity.
Before starting the asymptotic analysis of the functional (6.17) it is convenient to introduce the dimensionless co-ordinates
x

, S,
=
h
where S does not depend on h and has the diameter 1. Then h enters the
action functional explicitly through the components of the strain tensor ab
and the electric field Ea . In terms of , x the components of ab and Ea read
1
w(|) ,
h
1
1

.
3 = [w,x w $e.
. w + (1 + h )w| + $e. w| ],
2
h

.
33 = (1 + h )(w,x + w ) + h$e.
. (w,x w $e. w ),
1
E = | , E3 = ,x .
h

 =

6.2. ASYMPTOTIC ANALYSIS

217

We use the vertical bar preceding a Greek index to denote the differentiation
with respect to . The integral over S will be denoted by h.i.
Another form of the electric enthalpy. We now transform the electric
enthalpy density to another form more convenient for the asymptotic analysis. Note that among terms of W (ab , Ea ) the derivatives w| /h and w| /h
in  and 3 as well as E = | /h are the main ones in the asymptotic sense. Therefore we shall group principal terms and principal cross
terms containing  , 3 and E in the electric enthalpy. To this end let us
decompose the electric enthalpy (5.28) as follows
W = W + Wk ,

(6.20)

where
Wk = min max W.
 ,3 E

After some algebra we find that


1
Wk = (cK 233 2eK 33 E3 K E32 ),
2
1
+ c33
+ c3

W = (c
E
E
2 E
2e F 2e3 F
S F F ),

(6.21)

=  + k 33 t E3 ,
= 23 + r 33 s E3 ,
F = E + p 33 + q E3 .

(6.22)

where

The coefficients cK , eK , K and k , t , r , s , p , q are expressed through


the 3-D electroelastic moduli in the following way
cK = c3333
+ p e33
H
H ,

3
eK = e333
H p H ,

3
K = 33
H q H ,
H 3
q =
H ,

H 33
p =
eH ,

H
1

= (
H ) ,

c3333
= c3333 f c333 , e333
333 g c333 ,
H
H = e
e33
33 g c333 , a
a + g ea3 ,
H = e
H =

218

CHAPTER 6. PIEZOELECTRIC RODS


33
33 + g e33 ,
H =
g = s e33 ,

f = s c333 ,

g = s e3 ,

s = (
c33 )1 ,

(6.23)

a3
c
=
f c33
, c3a3 = c3a3
f
cE ,
E
E
a3
3a3
a3
a3
a3
e = e h cE , e = e g cE ,
3
a
,
, 33 = 33
a = a
S + h e
S + g e

f = b c33
, f
= b c3
, g
= b e ,
E
E
h = b e3 , r = f + g p , s = g g q ,

q ,
s g
p , t = h f
r + g
k = f f
3333

c3333
E
3a3

where b is the tensor inverse to c


in the following sense: b c
=
E
E
( )
.
Two-dimensional tensors of electroelastic moduli. The electroelastic
constants c
, c33
, c3
, e and
E
E
E
S satisfy the following symmetry
properties
c
= c
= c
= c
,
E
E
E
E
33
cE
= c33
,
E

e = e ,

c3
= c3
,
E
E

S = S .

These constants together with e3 , k , t , r , s , p , q , cK , eK and K


can be regarded as independent components of the two-dimensional tensors
of electroelastic moduli.
We now consider piezoelectric rods that are homogeneous over the cross
sections. The components of the 3-D moduli of such a rod, referred to the
basis {e , e}, should not depend on . When these tensors are referred
to the basis {t , t}, their components will in general depend on as evident from the relation (4.19). Therefore the components of the 2-D moduli
depend in general also on . It can be shown for homogeneous rods, however, that components of any two-dimensional electroelastic moduli, denoted
symbolically by A( , x), possess the property
h
A( , x) = A0 (x) + O( )A0 (x),
R
where A0 are their values evaluated at = 0. Therefore, when constructing
equations of vibrations of rods admiting the error h/R as compared with
unity, it can be assumed that A = A0 , i.e., the 2-D moduli of rods homogeneous over the cross sections are independent of the transverse co-ordinates
.

6.2. ASYMPTOTIC ANALYSIS

219

We note some special symmetry cases, where the independent electroelastic constants can be reduced in number.
Mirror planes perpendicular to the central line. If the electroelastic
properties are invariant under reflections relative to planes perpendicular to the central line, then the following relations
c3
= 0,
E

e3 = 0,

q = 0,

t = 0,

r = 0,

and
eK = e333
H ,

K = 33
H,

hold true.
n-fold rotation axis that is tangential to the central line. When n is
even, all 2-D tensors of odd rank vanish.
Transverse isotropy. When the electroelastic properties are invariant
under rotations about the tangent to the central line (model of a piezoceramic rod polarized along the central line with symmetry m), it
can be shown that all 2-D tensors of odd rank vanish; the tensor c
E
has the form
c
= a a + (a a + a a ),
E
and all the 2-D tensors of second rank are spherical.
From the definition of Wk one can also obtain the following interesting
formulae
k2
1
,
e
=

,
K
2
g33 (1 + k 2 )
sD
33 (1 + k )
1
(g33 )2
2
K = T
,
k
=
,
T
33 (1 + k 2 )
sD
33 33
1
da3
a3 b3
ab
= E , ea3
=
, ab
H
H = T d eH ,
s33
sE
33

cK =

c33
H

(6.24)

where the abbreviated indicial notation is used for the components of the
fourth and third rank tensors. Indeed, using the definition (2.49) for the
elastic enthalpy F (, D) we have
F ( ab , Da ) = min max[Da Ea ab ab + W (ab , Ea )].
ab

Ea

220

CHAPTER 6. PIEZOELECTRIC RODS

Making the substitutions = 0, 3 = 0 and D = 0 in this formula and


remembering the definition of Wk , we obtain
F (0, 0, 33 , 0, D3 ) = min max[D3 E3 33 33 + Wk (33 , E3 )].
33

E3

But according to (2.51)


1
1 T
33 2
33 3
3 2
F (0, 0, 33 , 0, D3 ) = sD
33 ( ) g33 D + 33 (D ) .
2
2
Comparison with (6.21)1 yields (6.24)1,2 . The remaining formulae then follow
from (6.23).
Asymptotic analysis of the action functional.
Case i) Unelectroded lateral surface. Similar to the derivation for piezoelectric
shells, we shall omit the first steps of the variational-asymptotic procedure
and make a change of unknown functions immediately. We seek the displacements w , w and the electric potential in the form
w = u (x, t) he (x, t) + hy ( , x, t),

w = u(x, t) he.
. (x, t) + hy( , x, t),
= (x, t) + h( , x, t),

(6.25)

where u , u, and are functions of x and t, being expressed through u


and u by (6.5). Without limiting generality we can force the functions y , y
and to satisfy the constraints
hy i = 0,

hyi = 0,

hy| ie = 0,
hi = 0.

(6.26)

According to (6.26) u , u and describe the mean displacements and electric potential of the rod, respectively, while corresponds to the mean rotation of the cross section. Equations (6.25) and (6.26) set up a one-to-one
correspondence between w , w, and the set of functions u , u, , , y , y,
and can be regarded as the change in the unknown functions {w , w, }
{u , u, , , y , y, }.
Asymptotic analysis enables one to determine the order of smallness of
y , y, . If these functions vanished, then the electroelastic state of a rod
would be characterized by the measures of elongation , bending and

6.2. ASYMPTOTIC ANALYSIS

221

twist as well as the 1-D electric field F = 0 as given by (6.4) and (6.6),
respectively. We introduce the following notations
p
 = h max + 2 , fF = max |F |,
x
x
x
q
q
= max y | y| + y | y| , = max | | .

 = max ||,
B

Consider a point x (0, L). The best constant l in the inequalities



fF

, h|a,x | , |F,x |
,
l
l
l

max |y,x | , max |y,x | , max |,x |


l
l
l
S
S
S
|,x |

is called the characteristic scale of change of the electroelastic state in the


longitudinal direction of the rod. We define the inner domain ci as a portion
of c(x) for which the following inequalities hold:
h? = h/R  1,

h?? = h/l  1.

(6.27)

We assume that the remaining portion cb is localized at the edges of the rod
and its width is of the order h (boundary layer). The functional (6.18) can
then be decomposed into the sum of two functionals, an inner one for which
an iteration process will be applied, and a boundary layer functional. It turns
out that the latter can be neglected in the first-order approximation. Therefore, the problem reduces to finding stationary points of the inner functional
that can be identified with the functional (6.18).
Keeping u , u, , fixed, we seek y , y, . Substituting (6.25) into the
action functional (6.18) and taking (6.19) into account, we neglect the time
rates of y , y in the kinetic energy. Proceeding with the estimations based
on the inequalities (6.27) similar to those provided in (4.35)-(4.37) we derive
the asymptotic formulae
 = y(|) , 23 = y| he ,
33 = + h ,
E = | , E3 = F.
This means that the partial derivatives of y , y and with respect to x do
not enter the action functional. Besides, the functions y , y and do not
enter the longitudinal energy. With (6.21), (6.22) the determination of these

222

CHAPTER 6. PIEZOELECTRIC RODS

functions reduces to the following minimax problem:


= h2

inf

sup h( , , F )i,

(6.28)

y ,y(6.26) (6.26)

1
( , , F ) = (c
+ c33
+ 2c3

E
E
E
2
2e F 2e3 F
S F F ),

(6.29)

where
= y(|) + k ( + h ) t F,
= y| he + r ( + h ) s F,
F = | + p ( + h ) + q F.

(6.30)

The functional (6.28) represents the transverse electric enthalpy, integrated


over the cross section of the rod. The variational problem (6.28) is called the
cross section problem.
Assume that the cross section problem (6.28) is solved and the functions
y , y and extremizing the functional (6.28) are found. In accordance with
the variational-asymptotic scheme we represent w , w and by (6.25) with
y , y, replaced by y , y, ,
but now the functions u , u, and should
be regarded as unknown functions. We substitute (6.25) into the action
functional (6.18) and integrate over the cross section. Within the first-order
approximation one can put g = 1. On the chosen field the average transverse
electric enthalpy takes the minimum value which is equal to , while the
average longitudinal electric enthalpy should be calculated with
33 = + h ,

E3 = F.

The average longitudinal electric enthalpy per unit length of the piezoelectric
rod is thus given by
k =

h2
hcK ( + h )2 2eK ( + h )F K F 2 i
2
1
= (cK |S| 2 + cK I 2eK |S|F K |S|F 2 ), (6.31)
2

where I = h4 h i are the moments of inertia of the cross section.


Since the kinetic energy density contains only w and w,
and since the
principal terms of the last two are the same as for elastic rods, the average
one-dimensional kinetic energy is given by
1
1
= |S|(u 2 + u u ) + Ip 2 ,
2
2

(6.32)

6.2. ASYMPTOTIC ANALYSIS

223

where Ip = I is the polar moment of inertia of the cross section. Comparing


(6.32) and (6.8), we see that = |S| and = Ip /|S|.
Thus, the problem of constructing the 1-D theory of vibrations of piezoelectric rods reduces to the solution of the cross section problem (6.28).
Case ii) Partially electroded lateral surface. In this case the electric potential
should satisfy the constraints (6.3) so that another change of the unknown
functions is more appropriate:
w = u (x, t) he (x, t) + hy ( , x, t),

w = u(x, t) he.
. (x, t) + hy( , x, t),
= h( , x, t).

(6.33)

We impose on the functions y , y and the following constraints


hy i = 0,

hyi = 0,

hy| ie = 0,
|c(i)
= (i) (t)/h,
e
(i)

(6.34)

(i)

where ce = { | h ce }. The last constraint in (6.34) is due to the


boundary conditions (6.3) on the electrodes.
Introduce the following notations
p
 = max ||,  = h max + 2 ,
x
x
q
q
|
|
= max y y| + y y| , = max | | .
B

We define the characteristic scale of change of the electroelastic state in the


longitudinal direction of the rod as the best constant l in the inequalities


|,x | , h|a,x | ,
l
l
(6.35)

max |y,x | , max |y,x | , max |,x |


l
l
l
S
S
S
We assume, as in case i), that h/l  1 everywhere in the rod except at a
boundary layer of width of the order h. Since the energy of the boundary
layer is small compared with that of the inner region, it will be neglected in
the first-order approximation.
The estimation based on (6.35) yields the asymptotic formulae
 = y(|) , 23 = y| he ,
33 = + h ,
E = | , E3 = 0.

(6.36)

224

CHAPTER 6. PIEZOELECTRIC RODS

With (6.33) and (6.36) the asymptotic analysis of the action functional (6.18)
can be done. Not going into the details of this analysis, which is similar to
that given in case i), we present the results below. Within the first-order
approximation the 3-D action functional (6.18) can be replaced by the 1-D
functional (6.13), where
1
1
= |S|(u 2 + u u ) + Ip 2 ,
2
2
1
= (cK |S| 2 + cK I ) + (, , ).
2

(6.37)

The average transverse electric enthalpy per unit length of the piezoelectric
rod (, , ) must be determined by solving the cross section problem
= h2

inf

sup h( , , F )i,

(6.38)

y ,y(6.34) (6.34)

where the function ( , , F ) is given by (6.29), with


= y(|) + k ( + h ),
= y| he + r ( + h ),
F = | + p ( + h ).

(6.39)

If the extremizers y , y and are found, the asymptotics of the 3-D electroelastic state of the rod can be restored according to the formulae (6.33) and
(6.36).
Problems
1. Derive the additive decomposition of the electric enthalpy (6.20)-(6.21)
and obtain the formulae (6.22) for the 2-D electroelastic moduli.
2. Follow the variational-asymptotic scheme to determine the set N and
to justify the change of unknown functions (6.25) in the problem i).
3. Perform a similar procedure to justify (6.33) in the problem ii).

6.3

Cross section problems

Unelectroded lateral surface. In this case the cross section problem


(6.28), (6.29) can be simplified considerably. Indeed, since the piezoelectric
rod is assumed to be homogeneous, the 2-D electroelastic moduli can be
regarded as independent of in the first-order approximation. Let us now

6.3. CROSS SECTION PROBLEMS

225

make in (6.29) a change of unknown functions {y , y, } {z , z, } as


follows
1
h i
y = k + t F a (
) + hz ,
2
|S|
y = r + s F + hz,
= p + q F + h,
is the area of the cross section S and
where |S|
a = h(k + k k ).
Due to (6.26) the new functions should satisfy the following constraints
hz i = 0,

hzi = 0,

hz| ie = 0,

h i = 0.

(6.40)

It is easy to check that the variational problem (6.28), (6.29) reduces to the
following problem
= h4

inf

sup h( , , F )i,

(6.41)

z ,z(6.40) (6.40)

where ( , , F ) is given by (6.29), but now


= z(|) , = z| e + r ,
F = | + p .
The parameters entering this variational problem are and . Since the
functional is quadratic, we have
1
= (A + 2B + C2 ).
2
Combining this with equation (6.31) we arrive at the formula (6.15).
The transverse 1-D electric enthalpy can be found explicitly under the
following assumptions: i) there exists in each point of the rod an even-fold
rotation axis parallel to the central line, ii) the cross section S is ellipse.
Due to the first assumption all the 2-D tensors of odd rank vanish, and the
functional (6.41) breaks into the sum of two functionals
1
h1 i = h4 hc
i,
2 E
and
1
h2 i = h4 hc33
2e3 F
S F F i,
2 E

(6.42)

226

CHAPTER 6. PIEZOELECTRIC RODS

where
= z(|) ,

= z| e ,

F = | .

The infimum of the first functional is obviously zero and is attained at z = 0.


Consider the second functional (6.42) and assume the cross section to be given
in the form
c 1,
where c are the components of a positive definite symmetric second-rank
tensor. Varying (6.42) we obtain for z and the Euler equations
c33
(z| e )| + e3 | = 0,
E
e3 (z| e )|
S | = 0,

(6.43)

and the corresponding natural boundary conditions


c33
(z| e ) + e3 | = 0,
E
e3 (z| e )
S | = 0,

(6.44)

c = 1, where is the unit outward normal to


at the boundary S:

= c /|c |. Similar to the elastic rods of elliptical cross sections,


S:
the solution of (6.43) and (6.44) is sought so as to satisfy the equations
c33
(z| e ) + e3 | = ae c ,
E

e3 (z| e )
S | = be c ,

(6.45)

with a, b still unknown constants. From (6.45) we obtain


= z| e = A ,
F = | = B ,

(6.46)

where the tensors A and B are given by

3 1
S e c ),
A = c1
33 (ae c + be
3 1
3 1
B = 1
c33 e c + b(e3 c1
S e c e c )],
S [ae
33 e

and
3 3
c33 = c33
+ 1
e .
E
S e

Equations (6.45) have a solution if and only if


e A = 2,

e B = 0.

6.3. CROSS SECTION PROBLEMS

227

These conditions determine uniquely the constants a, b.


Substituting the solution (6.46) into the functional (6.42), we obtain
1
= C2 ,
2
where

C = (c33
A A + 2e3 A B
E
S B B )I ,

and I = h4 h i are the moments of inertia of the cross section. In the


special case of piezoceramic rods polarized along their central line, it can
easily be shown that
1 1
C = 4cE
) ] ,
55 [(I

1313
(cE
55 = cE ).

Summing up, for the piezoelectric rod of the elliptical cross section that
possesses the even-fold symmetry axes along the central line, we obtain the
expression for the 1-D electric enthalpy in the form
1
= (cK |S| 2 + cK I + C2 2eK |S|F K |S|F 2 ).
2
Partially electroded lateral surface. This case differs from the previous
one by the greater variety of types of piezoactive vibrations. Here we need
to solve the variational problem (6.38). It can be simplified by making the
change of unknown functions {y , y, } {z , z, } according to
h i
1
y = k a (
) + z ,
2
|S|
y = r + z, =
where
a = (k + k k ).
The variational problem (6.38) then reduces to
= h2

inf

sup h( , , F )i,

(6.47)

z ,z(6.48) (6.48)

under the constraints


hz i = 0,

hzi = 0, hz| ie = 0,
|ce(i) = (i) (t)/h,

(6.48)

228

CHAPTER 6. PIEZOELECTRIC RODS

where the function ( , , F ) has the same form as (6.29), with


= z(|) , = z| he + r h ,
F = | + p ( + h ).
Note that among the constraints (6.48) the only ones being imposed on are
essential, since the functional (6.47) is invariant with respect to the transformations z z + c and z z + c + e , where c, c , are constants.
We further assume that the cross section of the rod is centrally symmetric,
then S,
and if c(i)
which means that if S,
e is covered by an
(i)
(i)

electrode, then ce = { | ce } is covered by another one too. For


functions defined in centrally symmetric domains the concept of evenness and
oddity can be introduced, and each function can be uniquely decomposed into
the sum of odd and even ones (these will be marked with one and two primes,
respectively). For rods with centrally symmetric cross sections one can thus
show that the variational problem (6.47) breaks into two problems
= 1 (, (i) (i) ) + 2 ( , , (i) + (i) ).

(6.49)

The part 1 (, (i) (i) ) is the solution of the variational problem


1 (, (i) (i) ) = h2 inf
sup h( , , F )i,
0
0
z ,z 0 (6.51)

(6.50)

under the constraints


1
0 |ce(i) = ((i) (t) (i) (t))/h,
2

i = 1, . . . , n,

(6.51)

where the function ( , , F ) has the same form as (6.29), with


0
= z(|)
,

0
= z|
,

F = 0| + p .

Analogously, 2 ( , , (i) + (i) ) is the solution of the variational problem


2 ( , , (i) + (i) ) = h2 inf
00 00

sup h( , , F )i,

z ,z 00 (6.53)

(6.52)

under the constraints


1
00 |ce(i) = ((i) (t) + (i) (t))/h,
2

i = 1, . . . , n,

with
00
= z(|)
,

00
= z|
he + r h ,

F = 00| + p h .

(6.53)

6.3. CROSS SECTION PROBLEMS

229

Figure 6.1: Cross section of a piezoelectric rod.


Consequently, if (i) (t) = (i) (t) for all i, then 1 vanishes and, alternatively, if (i) (t) = (i) (t) for all i, then the cross terms between , and
(i) in 2 vanish. Due to the arbitrariness of the electrode arrangement, it
is difficult to solve the variational problems (6.50) and (6.52) in the general
case. Therefore we shall now analyze some concrete examples.
Example 1. Piezoelectric rods with rectangular cross sections. The dimensionless width of the cross section is denoted by b = b/h. On the sides
1 = b/2 there are two electrodes, and the values of the electric potential
0 (t)/2 are prescribed on them (see Figure 6.1). From the previous consideration it is clear that only 1 possesses the cross term between and
0 (t). In order to find it we have to solve the problem (6.50). The latter will
be transformed to the dual variational problems according to the standard
procedure [6, 13]. Thus, 1 can be found from the minimization problem
2

1 = h

inf [h( , , Q ) Q p i +

0 ,z 0 ,Q
z

Z
2
X
i
i=1

(i)
ce

Q ds],

(6.54)

under the constraints


Q| = 0,

Q = 0 on cd ,

where 1,2 = 0 /2 and ( , , Q ) is the Legendre transformation of


the quadratic form with respect to F
( , , Q ) = max[F Q + ( , , F )].
F

(6.55)

230

CHAPTER 6. PIEZOELECTRIC RODS

On the other side, 1 can also be determined by solving the maximization


problem
1 = h2

sup h? ( , , F )i,

(6.56)

, ,0

under the constraints

| = 0,

| = 0,

= 0, = 0 on S,
0
|c(i)
= i /h,
e
where ? ( , , F ) is the Legendre transformation of the quadratic form
with respect to and
? ( , , F ) = max [ + ( , , F )].
,

(6.57)

These dual variational problems enable one to establish the upper and lower
bounds for the function 1 .
Let us choose the trial functions in (6.54) as follows
1
z0 = a ,
2

z 0 = a ,

Q1 = a,

Q2 = 0.

Substituting this into the functional (6.54), we obtain the inequality


1 |S|[(a , a , a, 0) ap1 + a

0
].
b

Minimization of the right-hand side of this inequality with respect to a , a


and a leads to the upper bound for 1 . Taking into account that
1 H 2
min (a , a , a, 0) = 11
a,
2

a ,a

H
which follows from the definition (6.55) and the formula (6.29), with
the

inverse tensor to H : H = , we obtain

|S|
H
(p1 0 /b)2 /11
.
2

The lower bound can be established, if we substitute in (6.56)


= = 0,

0 =

0 1
1
+ a cos 2 ,
b
b

(6.58)

6.3. CROSS SECTION PROBLEMS

231

and then minimize the right-hand side of the obtained inequality with respect
to a. Recalling the following useful relation
1
F F ,
? (0, 0, F ) =
2 H
which follows from the definition (6.57) and the formula (6.29), we find that
1

|S|
[ (p + G )(p + G ) + A2 /B],
2 H

(6.59)

where
G1 =
A=

0
,
b

G2 = 0,

2 22
0
[H p2 + 12
)],
H (p1

B=

22
11 2
H
+ H 2 .
2
24b

The upper and lower bounds (6.58) and (6.59) coincide, when 12
H = 0, p2 =
0.2 In this case
1 =

|S| 11 2 2
0
0
H [p1 2p1
+ ( )2 ].
2
b
b

Combining this with (6.37) and (6.49) we get


0
1
= (c|S| 2 + cK I + 2e|S| ) + 2 ( , ),
2
b

(6.60)

where c, e are given by


c = c33
H =

1
,
sE
33

e = e13
H =

d13
.
sE
33

(6.61)

Besides, we find that the component D1 of the electric induction is equal to


D1 = e 11
H

0
.
b

(6.62)

This formula is needed for the calculation of the antiresonant frequencies.


Note that the formulae (6.60)-(6.62) are not valid in the general case of
anisotropy, as is seen from the inequality (6.59). Particularly, by putting
0 = 0 therein, one can see that c > c33
H when A 6= 0.
Example 2. Piezoelectric rods with rectangular cross sections. Its opposite
sides 1 = b/2 are covered by two pairs of electrodes: at 1 = b/2 we have
= 0 /2h for 2 < 0 and = 0 /2h for 2 > 0; at 1 = b/2 we have

232

CHAPTER 6. PIEZOELECTRIC RODS

Figure 6.2: Cross section of a piezoelectric rod.


= 0 /2h for 2 < 0 and = 0 /2h for 2 > 0 (see Figure 6.2). Due
to this electrode arrangement and the boundary conditions for the electric
potential 1 = 0. For simplicity let us assume that the rod possesses the
mirror planes perpendicular to the central line and, additionally, the evenfold rotation axis in the direction 1 . Then c3
= e3 = p2 = 12
H = 0
E
and
2 = 3 ( , 0 ) + 4 ().
Thus, only the part 3 ( , 0 ) contains the cross terms between the measures
of bending and 0 .
The variational problem for 3 ( , 0 ) can be simplified by making the
change of unknown function 00 according to
00 =

h
p1 1 [( 1 )2 b2 /4] + .
2

One can then see that 3 ( , 0 ) does not depend on 1 and can be found
by solving the following minimax problem
1
hcE 2e F
S F F i,
2
(6.63)

3 = h2 inf
sup
00
z

under the constraints


0
2h
0
=
2h

| 1 =b/2 =
| 1 =b/2
2

for 2 0,
for 2 0,

This is the case for the rod with the even-fold rotation axis parallel to 1 .

(6.63)

6.3. CROSS SECTION PROBLEMS

233

where
00
,
= z(|)

F1 = |1 + p1 h2 1 ,

F2 = |2 .

Passing to the dual maximization problem, one can easily prove that
1
3 h2 sup h
F F i
2 H
(6.63)
|S|[
where

1 11
0
911 0
H (p1 h2 + 3 )2 (A H )( )2 ], (6.64)
24
b
24
b

1
2

A = sup h 11
H | | i/(b(0 /b) ).
2
(6.63)

From the other side, changing to the dual minimization problem and performing the same procedure as in the previous case, we obtain the upper
bound
3 |S|

1 11
0
H (p1 h2 + 3 )2 .
24
b

(6.65)

From (6.64) and (6.65) follows


3 = |S|

1 11
0
0
[H (p1 h2 + 3 )2 + c( )2 ]
24
b
b

This result is remarkable because of the fact that the solution of the cross
section problem is not known, and nevertheless 3 was found exactly up to
the unimportant constant.
Problems
1. Consider piezoceramic rods polarized along their central lines, with
unelectroded lateral surfaces. Establish the following upper bound for
the torsional rigidity
4cE
C 155
(I )
for arbitrary cross sections.
2. Show that for the piezoceramic rod of rectangular cross section polarized along the axis 1 , with the electrodes as in example 1, 3 = 0.

3. Find an example where cross terms between and i occur in .

234

6.4

CHAPTER 6. PIEZOELECTRIC RODS

Frequency spectra

Straight and untwisted end-electroded rod. If the lateral surface of


the rod is not covered by electrodes, then the 1-D enthalpy is given by the
formula (6.15). It is easy to see that for the naturally straight and untwisted
rod the longitudinal and flexural-torsional vibrations are uncoupled, and only
the former is piezoelectrically active. The ends of the rod are covered by a
pair of electrodes with = 0 /2eit specified on them. To determine the
frequency spectra of the longitudinal vibrations we write down the equations
of motion

u = (cK u0 + eK 0 )0 ,
(eK u0 K 0 )0 = 0,

(6.66)

which are subject to the boundary conditions at x = 0, L


cK u0 + eK 0 = 0,
0
= eit .
2

(6.67)

Integration of the second equation of (6.66) yields


eK u0 K 0 = d0 eit ,

(6.68)

where d0 is a constant. Solving this equation with respect to 0 and substituting the result into the first equation, we obtain

u = cK u00 ,

cK = (cK + e2K /K ).

(6.69)

1/2
Using the relations (6.24) one can check that a = (
cK /)1/2 = (1/sD
33 )
corresponds to the velocity of the longitudinal waves in rods. The harmonic
solution to (6.69) gives longitudinal displacement of the form

u = (b1 sin

x
x it
+ b2 cos
)e .
a
a

At the free ends of the rod the tension is zero (boundary condition (6.67)1 ),
which can be used to evaluate b1 and b2 , giving
u=

a
x
L
x it
g33 d0 (sin
tan
cos
)e .

a
2a
a

(6.70)

The extension u0 , obtained from (6.70), is substituted into (6.68), which is


then integrated to obtain the voltage across the rod
Z L
L
it
T a 2
0 e =
0 dx = 33
[ k 2 tan
L(1 + k 2 )]d0 eit ,

2a
0

6.4. FREQUENCY SPECTRA

235

2
T
where k 2 = e2 /
cK K = g33
/sD
33 33 is the piezoelectric coupling factor. From
this equation one can determine the remaining unknown coefficient d0

d0 =

2
T
L[ kf
33

0
,
tan f (1 + k 2 )]

where f = L/2a is the dimensionless frequency. The resonant frequencies


are determined from the condition d0 = giving
1
1 + k2
tan fr =
.
fr
k2
The antiresonant frequencies are determined from the condition that the
electric current is zero or, equivalently, d0 = 0. This leads to
fa =

(2n 1),
2

n = 1, 2, . . . .

Straight and untwisted side-electroded rod. Consider a rod of rectangular cross section, with electrodes on the lateral surface as in example 1,
Section 6.3. We assume also that 12
H = p1 = 0, and the 1-D electric enthalpy
is given by (6.60). Again, for the naturally straight and untwisted rod the
longitudinal and flexural-torsional vibrations are uncoupled, and only the
former is piezoelectrically active. The equation of motion takes the simple
form

u = cu00 .
For 0 (t) = 0 eit , we seek u = ueit as the harmonic solution of this equation
with the frequency giving
u = (b1 sin

x
x
+ b2 cos
),
a

p
where a
= c/ is the velocity of the longitudinal waves in rods with the
pair of electrodes on the lateral surface. Since the rod ends are free we have
to satisfy the following boundary conditions at x = 0, L
c
u0 + e

0
= 0,
b

yielding
b1 =

a
e 0
,
c b

b2 =

a
e 0
L
tan
.
c b
2
a

236

CHAPTER 6. PIEZOELECTRIC RODS

Having found the displacement u we now determine the electric induction


D1 according to (6.62) and then integrate along the rod to obtain the total
charge on the electrode
Z L
a
e2 0
L
0
D1 dx = (
2 tan
11
L)eit .
H

c
b
2
a
b
0
The resonant frequencies are determined from the condition that the total
charge becomes infinite giving
tan fr = fr =

(2n 1),
2

n = 1, 2, . . . ,

where f = L/2
a is the dimensionless frequency. The antiresonant frequencies are determined from the condition that the total charge is zero. This
leads to the equation
k2 1
1
tan fa = 13 2 ,
fa
k13
2
T
3
where k13
= d213 /sE
33 11 is the piezoelectric coupling factor.

Figure 6.3: A piezoceramic open circular ring.

Piezoceramic side-electroded open circular ring. Consider a thin


piezoceramic open circular ring of rectangular cross section. The radius of
the central line is denoted by R, the width and the thickness of the cross
section by b and h, respectively. The rod is polarized along its width and
3

In acoustics [9] the direction of polarization is normally denoted by x3 and the coordinate along the rod axis by x1 , so the piezoelectric coupling factor in this case should
2
T
be k31
= d231 /sE
11 33 .

6.4. FREQUENCY SPECTRA

237

electroded on its outer and inner faces (see Figure 6.3). From the previous
section it follows that the 1-D enthalpy of such a rod should have the form
0
1
+ C2 ),
= (c|S| 2 + cK I + 2e|S|
2
b
where 0 (t)/2 are the given values of the electric potential on the electrodes.
Thus, in this case the plane and flexural-torsional vibrations turn out to be
uncoupled, and only the former is piezoelectrically active. Choosing the triad
t , t as given by (4.94) and deriving the equations of motion describing the
plane vibrations in a similar manner as for the elastic ring, we obtain
1
1 00
1 0
u1 ) cK I11 (u000
u ),
1
R
R
R
1 000
1
1
0
|S|
u1 = |S|[c(u0 u1 ) + e ] + cK I11 (u0000
u ).
1
R
R
b
R
Introducing the dimensionless quantities
r
t c
x
cK I11
=
, = , ? =
.
R
R
|S|cR2
|S|
u = c|S|(u00

(6.71)

we rewrite (6.71) as follows (cf. equations (4.96))


00
u| = u00 u01 ? (u000
1 u ),
e
000
u1| = u0 u1 + 0 + ? (u0000
(6.72)
1 u ),
c
where 0 = R0 /b, and where the prime denotes the derivative with respect
to . Choosing the co-ordinate such that the rod occupies the segment
|| 0 in its undeformed state, we write down the traction free boundary
conditions at = 0 :
e
u0 u1 + 0 = 0,
c
00
(6.73)
u1 + u0 = 0,
00
u000
1 + u = 0.

Assuming the voltage 0 ( ) a harmonic function of : 0 ( ) = 0 ei , we


seek solutions of (6.72), (6.73) in the form
u = uei ,

u1 = u1 ei .

(6.74)

Eliminating the common factor ei in (6.72), one can show that its symmetric solution is given by
u =

3
X
i=1

ai sin i ,

u1 =

X
e 0
+
ai i cos i ,
c 1 2 i=1

(6.75)

238

CHAPTER 6. PIEZOELECTRIC RODS

where 2i are the roots of the cubic equation


(2 + ? 2 2 )(1 + ? 4 2 ) 2 (1 + ? 2 )2 = 0,

(6.76)

and i are expressed through i in the following way


i =

(1 + ? )2i 2
.
i (1 + ? 2i )

For each there are three different roots of equation (6.75) up to the sign of
, which are calculated for ? = 0.001 and plotted in Figure 6.4. In a small
range near the zero frequency there are three different real roots, then up to
the frequency two complex conjugate roots and one real root, in the range
( , c ) two different imaginary roots and one real root, and, finally, above
the cut-off frequency c one imaginary root and two different real roots.

Figure 6.4: Roots of equation (6.76) for ? = 0.001.


We now substitute the general solution (6.75) into the boundary conditions (6.73) to obtain the following system of linear equations for the determination of the coefficients ai
3
X

Cij aj = bi ,

i = 1, 2, 3.

j=1

The elements of the matrix Cij are given by


C1j = (j j ) cos j 0 ,
C2j = (j 2j + j ) cos j 0 ,
C3j = (j 3j 2j ) sin j 0 ,

6.4. FREQUENCY SPECTRA

239

and bi by
b1 =

e 0 2
,
c 1 2

b2 = b3 = 0.

Figure 6.5: Resonant frequencies of piezoceramic side-electroded circular


rings with free edges (? = 0.001, k13 = 0.344).
In order to calculate the amplitude of the total charge on one of the electrodes, we should know the amplitude of the electric induction D1 . According
to (6.62)
3
11
2 X
e2
1
1 = H 0 [ e
D
a

)
cos

1],
(6.77)
i i
i
i
11
11
R cH i=1
cH 1 2
where a
i are the solution of the following system
3
X

Cij a
j = bi ,

i = 1, 2, 3,

j=1

b1 =

2
,
1 2

b2 = b3 = 0.

Integration of (6.77) gives the amplitude of the total charge on one of the
electrodes
Z 0
3
e2 X i i
e2
1
1
11

D Rd = 2H 0 [ 11
+ 1)0 ].
a
i
sin i 0 ( 11
cH i=1
i
cH 1 2
0
Thus, the resonant frequencies correspond to the roots of the equation
det Cij = 0,

240

CHAPTER 6. PIEZOELECTRIC RODS

while the antiresonant frequencies should be found from the condition of


2
2
vanishing charges on electrodes. Noting that e2 /c11
H = k13 /(1 k13 ), we
obtain the following equation determining the antiresonant frequencies
3

2
X i i
1 k13
1
)
=
sin i 0 .
+
a
i
(
0
2
1 2
k13
i
i=1

Taking for example ? = 0.001, k13 = 0.344 (which is the piezoelectric


coupling factor of the PZT-5 ceramics), the first three resonant frequencies
in the range (0.14, 1.14) as functions of the parameter 0 are numerically
evaluated and the results are plotted in Figure 6.5. Within this range the
antiresonant frequencies differ very little from the resonant ones, and we do
not show them in the figure.
Problems
1. Determine the electromechanical coupling coefficient in the three cases
considered above.
2. Determine the resonant and antiresonant frequencies of the straight and
untwisted rod polarized along its width, whose faces are electroded in
the region x (L0 , L L0 ).
3. Determine the resonant and antiresonant frequencies of the piezoceramic side-electroded circular ring clamped at its ends.

6.5

Longitudinal impact

In this section the problem of the longitudinal impact of two piezoceramic


straight rods polarized along their central lines is solved.
Problem statement. Consider two piezoceramic straight rods of the same
material, length and cross section. Their free ends are covered by electrodes
that are short-circuited so that the electric potential vanishes there. In the
middle of the first rod an electrode is installed in the cross section. Let
the second rod move towards the first one that is at rest with the velocity
v0 . At t = 0 the two rods come into contact. The problem is to determine
the electroelastic state of the rods as well as the impact interval (the time
interval after which the rods separate). Let us choose the co-ordinate x
along the central line of the rods so that at t = 0 the first one is located

6.5. LONGITUDINAL IMPACT

241

at 0 < x < l. Since the rods are straight and polarized longitudinally, their
motion in the longitudinal direction is described by

u = cK u00 + eK 00 ,
eK u00 K 00 = 0,

(6.78)

for l < x < l. At the free ends of the rods the traction free boundary
conditions as well as the conditions of vanishing electric potential should be
fulfilled
cK u0 + eK 0 = 0, = 0,
(6.79)
at x = l. At the ends of the rods that are in contact the continuity
conditions of the displacement, electric potential, stress and electric induction
are posed
u(0) = u(+0), (0) = (+0),
(cK u + eK 0 )|x=0 = (cK u0 + eK 0 )|x=+0 ,
(eK u0 K 0 )|x=0 = (eK u0 K 0 )|x=+0 ,
0

(6.80)

where u(0) denote the limiting values of u as x approaches 0 from the right
and the left, respectively. Besides, we admit solutions having discontinuous
derivatives. Then at the point of discontinuity the continuity conditions
of displacement, electric potential and induction, as well as the following
dynamic condition
V [[u]]
= [[cK u0 + eK 0 ]],

(6.81)

should be fulfilled, where V is the velocity of the discontinuity front and


[[A]] = A+ A . These jump conditions are the consequences of the variational principle in dynamics (Section 6.1) and can be derived in a standard
way. Finally, the problem statement is completed by specifying the initial
conditions, which read
u|t=0 = 0 for l x l,
u|
t=0 = v0 for l x < 0,
u|
t=0 = 0 for 0 < x l.

(6.82)

Solution by the method of characteristics. In order to solve the problem


(6.78)-(6.82) we first express through u. From (6.78)2 we have
=

eK
u + C1 x + C2 .
K

242

CHAPTER 6. PIEZOELECTRIC RODS

The unknown constants C1 , C2 are determined from the second condition of


(6.79) yielding
C1 =

1
eK
[u(l) u(l)] ,
K
2l

C2 =

1 eK
[u(l) + u(l)].
2 K

Thus, the electric potential is given by


=

eK
1 eK
1 eK
u
[u(l) u(l)]x
[u(l) + u(l)].
K
2l K
2 K

(6.83)

Substitute this into (6.78)-(6.82) to obtain the equation

u = cK u00 ,

cK = cK + e2K /K ,

(6.84)

together with the following conditions


e2K
[u(l) u(l)] = 0 at x = l,
2lK
u(0) = u(+0), cK u0 |x=0 = cK u0 |x=+0 ,
(6.85)
= 0, u|
t=0 = v0 , l x < 0, u|
t=0 = 0, 0 < x l,
cK u0

u|t=0

and the jump condition on the front of discontinuity


V [[u]]
= [[
cK u0 ]].

(6.86)

The problem (6.84)-(6.86) resembles the standard problem of impact of two


elastic rods. The only difference between them is the non-local boundary
condition (6.85)1 , which is due to the long-range electric field. The problem
(6.84)-(6.86) can be solved by the method of characteristics. On Figure
p 6.6
the plane (x, t) with the system of characteristics x at = const, a = cK /
is shown. We first determine the function
u(l, t) u(l, t) = f (t).
For t l/a we find the following invariants
u au0 = 0 at x = l,
u + au0 = v0 at x = l.

(6.87)

Replace in equations (6.87) u0 |x=l = (e2K /2l


cK K )[u(l) u(l)] according
to (6.85)1 . Now subtracting from the first of (6.87) its second one, we get
ak 2
[u(l) u(l)] = v0 ,
l
[u(l) u(l)]|t=0 = 0,

[u(l)
u(l)]

6.5. LONGITUDINAL IMPACT

243

Figure 6.6: System of characteristics x at = const.


cK K is the piezoelectric coupling factor. Consequently, for
where k 2 = e2K /
t l/a
f (t) = u(l) u(l) =

a 2
lv0
(1 e l k t ).
2
ak

(6.88)

Analogously, for l/a t 2l/a we have


u au0 = v0 at x = l,
u + au0 = 0 at x = l.
Therefore
ak 2
f
f = v0 ,
l

f |t=l/a =

lv0
2
(1 ek ).
2
ak

(6.89)

It is easy to compute the solution of (6.89)


f=

a 2
lv0
2 a 2
(1 e l k t + 2ek e l k t ).
2
ak

(6.90)

The values of f in (6.88) and (6.90) are already enough for our aim. Consider
now the equation(6.84). Its solution reads
u = F1 (at x) + F2 (at + x).
It follows from (6.85)1 that u0 |x=l = u0 |x=l , therefore
F10 (at + l) + F20 (at l) = F10 (at l) + F20 (at + l).

(6.91)

244

CHAPTER 6. PIEZOELECTRIC RODS

This means that the function g 0 () = F10 () + F20 () is a periodic function


with the period 2l. Let us determine g 0 () on the interval (l, l). From the
initial conditions (6.85)3 we find that
F10 (x) + F20 (x) = 0 for l x x,
(
v0 l x < 0,
a(F10 (x) + F20 (x)) =
0 0 < x l.
From here we find the values of functions F10 () and F20 () on the interval
(l, l)
(
(
v
/2a
0
<

<
l,
0
0 < < l,
0
F10 () =
, F20 () =
0
l < < 0,
v0 /2a l < < 0.
Thus, on the interval (l, l) g 0 () = v0 /2a. Due to its periodicity g 0 () =
v0 /2a for all > l. Now the solution can be represented in the following
form
u = F (at x) F (at + x) +

v0
(at + x),
2a

(6.92)

where F () = F1 (). The solution is fully determined if the function F () is


known for > l. Above F 0 () has been found in the interval l < < l.
Let us compute it for > l. From the boundary condition (6.85)1 at the
right end follows
F 0 (at l) F 0 (at + l) +

k2
v0
= f (t).
2a
2l

Based on this equation of continuation we can compute F 0 (at+l) successively


on the intervals 0 at l, l at 2l, 2l at 3l, . . ., provided f (t) and
F 0 (at l) are known. The straightforward computations show that
v0 k2 (l)
el
,
2a

(6.93)

k2
k2
k2
v0
(2 e l (3l) + e l (l) 2e l (2l) ),
2a

(6.94)

F 0 () =
for l 2l, and
F 0 () =

for 2l 3l. This procedure can be continued, but we need only the value
of F 0 () on the interval (l, 3l).

6.5. LONGITUDINAL IMPACT

245

Impact interval and electroelastic state. The time interval after which
the rods separate is determined from the condition that u0 (+0, t) changes its
sign. According to the solution (6.92)
u0 (+0, t) =

v0
2F 0 (at).
2a

For 0 at l
u0 (+0, t) =

v0
< 0.
2a

Further, for l at 2l
u0 (+0, t) =

v0
v0 k2
e l (atl) < 0.
2a
a

For 2l at 3l
u0 (+0, t) =
Letting t

2l
a

k2
k2
k2
v0
[3 + 2e l (at3l) 2e l (atl) + 4e l (at2l) ].
2a

(6.95)

+ 0, we obtain the limit


u0 (+0,

2l
v0
2
2
+ 0) = [1 + 2ek 2ek ].
a
2a

One can check directly that this expression is always positive, provided
0 < k 2 < 0.24,

or |k| < 0.49.

(6.96)

This means that for a small coupling coefficient k 2 = e2K /


cK K satisfying the
inequality (6.96) the impact interval T is determined by
T =

2l
,
a

(6.97)

exactly as for elastic rods. The difference is that the kinetic energy of the
second rod will not be fully transmitted to the first one, and after their
separation they both continue to vibrate.
We now compute the difference of the electric potential on the electrodes
of the first rod during the impact. According to (6.83)
l
eK
l
1
1
( ) (l) =
{u( ) [u(l) u(l)] [u(l) + u(l)]}.
2
K
2
4
2

(6.98)

We find u(l/2) and u(l) + u(l) for 0 t 2l/a. From (6.92) we have
l
l
l
v0
l
u( ) = F (at ) F (at + ) + (at + ).
2
2
2
2a
2

246

CHAPTER 6. PIEZOELECTRIC RODS

Thus, the problem reduces to the computation of F (), which can be done by
the successive integration of F 0 (). For l 0 F () = 0. For 0 l
F () = v0 /2a. For l 2l
Z
v0 k2 (l)
v0 l
v0 l
1 k2
1
F () =
+
el
d =
(1 + 2 e l (l) 2 ).
2a
2a
k
k
l 2a
Similarly, for 2l 3l
l 2
l k2
v0
l
v0
l k2
(l + 2 ek 2 ) + [2 2 e l (3l) + 2 e l (l)
2a
k
k
2a
k
k
2l k2 (2l)
l k2
l k2 2l
2e l
4l + 2 e
2e 2]
k
k
k
k

F () =

For the sum u(l) + u(l) the following formula


u(l) + u(l) = v0 t
holds true for 0 t 2l. Now we can determine the voltage in (6.98). For
0 t l/2a we have
a 2
1
eK 1 v0 l
l
k t
l
) v0 t].
(1

e
( ) (l) =
[
2
2
K 4 ak
2

For l/2a t l/a we have


a 2
eK v0 l 1 v0 l
l
[

(1 e l k t )].
( ) (l) =
2
2
K 4a
4 ak

The maximum of the voltage can also be obtained. One can check that
|( 2l ) (l)| is attained at t = l/2a (the quarter of the impact time) and is
equal to
eK v0 l
1
l
2
|1 + 2 (1 ek /2 )|
|( ) (l)| =
2
K 4a
k
eK v0 l
=
+ O(k 4 ).
K 8a

(6.99)

The solution of this problem enables us to propose an alternative method


of experimental measurement for the piezoelectric moduli of ceramics. We
first measure l, and K by the standard method. Then, measuring the
impact interval T , we can determine a and cK from (6.97). Finally, measuring
the maximum of the voltage, we find eK and k 2 from (6.99). This method
differs from the experimental measurement of piezoelectric moduli based on
the resonance and antiresonance and can complement it as a test procedure.

6.5. LONGITUDINAL IMPACT

247

Problems
1. Determine the electroelastic state of the rods after the impact interval
T.
2. Solve the impact problem of the piezoceramic rod fixed at one end and
struck longitudinally at the other.
3. Do the same for the piezoceramic rod free at one end and struck longitudinally at the other.

248

CHAPTER 6. PIEZOELECTRIC RODS

Part II
High-frequency vibrations

249

Chapter 7
Elastic shells
7.1

Two-dimensional equations

Variational principle. Consider a shell with a middle surface S and a


thickness h as described in Section 3.1, and assume that it is made of a
homogeneous, isotropic elastic material. In Chapter 3 we gave prominence
to the consideration of low-frequency vibrations of thin shells. The problem
now is to construct a two-dimensional theory of shells that enables one to
describe their vibrations in a wider range of frequencies. Taking the variational approach as the basis, we postulate that the vibrations of the shell
in the absence of external body forces and surface tractions should occur in
accordance with the following variational principle
Z

t1

Z
( ) da dt = 0,

J =
t0

(7.1)

where and are the 2-D kinetic and strain energy density, respectively.
The functional (7.1) is called, as before, the 2-D action functional.
In the classical theory of low-frequency vibrations of elastic shells the 2D action functional (3.9) depends only on the three functions u (x , t) and
u(x , t) describing the mean displacements of the shell. These functions will
subsequently be called external degrees of freedom, since they characterize
fully the kinematics of the shell middle surface. It is natural to assume that,
as the frequency increases, some internal degrees of freedom corresponding
to branches of the shell thickness vibrations will be excited and become more
and more involved, so that they should be included as unknown functions
in the 2-D kinetic and strain energy densities. Thus, the construction of the
2-D shell theory, from this point of view, reduces to the problem of finding
i) the list of external and most essential internal degrees of freedom, and ii)
251

252

CHAPTER 7. ELASTIC SHELLS

the 2-D kinetic and strain energy densities depending on these degrees of
freedom. While doing so we require i) the linearity, ii) the hyperbolicity, and
iii) the asymptotical exactness of the 2-D Euler equations obtained from the
variational principle (7.1). By asymptotical exactness we mean the exactness
of the 2-D equations in the low- and high-frequency long-wave range up to
terms of the order h/R and h/L of smallness compared with 1, where R is
the characteristic radius of curvature of the shell middle surface and l the
characteristic wavelength in the longitudinal directions.
The essence of the proposed theory of high-frequency vibrations of elastic
shells can be expressed by the following formulae
=

h
(u 2 + u 2 + h2 2 + 2 + v 2 ),
2

(7.2)

h
h2
2
[s1 (A )2 + 2A A + (( )2 + ) + s2 ,
2
6
2

+ s3 (v;
) + 2v (;) v(;) + 2r1 h1 A + 2r2 h1 v;
+ 2r3 h1 v ,

+ f1 ( + )( + ) + f2 h2 2 + f3 h2 v v ], (7.3)
where = u, + b u , and the concise notation X2 = a X X is used
to denote the squared magnitude of the 2-D vector X. In (7.2)-(7.3) the
2-D densities of the kinetic and strain energy depend on the three functions
u , u corresponding to the external degrees of freedom and five additional
functions , and v, which represent the most essential internal degrees of
freedom within the range of frequencies of interest, where
A = u(;) b u,
= (;) + b( $) ,
1
$ = (u, u, ).
2
For the isotropic material the coefficients in (7.3) are found to be
 2
1 2
=
, f1 = ( 2 a + h2 s
4 ),
2 24
 2
f2 =
+ h2 s5 , f3 = (2)2 a + h2 s
4 ,
e
2 2
4
16
r1 =
, r2 = 2 , r3 = ,
2
e
3e
r3
2
8

1 2
s1 = 2 + 2 2 , =
, e=
,
e
1
2 2

(7.4)

(7.5)

7.1. TWO-DIMENSIONAL EQUATIONS

253

16(1 + 2e2 )2
1
16e cot(/2e)
8 2

,
+ 2 2
s2 = 2 +
e

9 (4e 1)e2 2 e2
16(1 + 2e2 )2
8e tan(e)
s3 = 1
,
2 2

9 (4e 1)e2
2

s
+ 6Hb 2Ka ,
4 = (3H K)a
s5 = H 2 (1/e2 16) + K(1/e2 8),

while is the mass density, the shear modulus and Poissons ratio.
Let the shell be referred to the curvilinear co-ordinates system {x , x} as
described in Section 3.2. In order to clarify the geometrical meaning of the
functions , and v we point out the following approximate formulae that
restore the 3-D displacement field of the shell from the functions u , u, ,
and v
w = u + a(x),
w = u x + h( + )q(x) + v p(x),
where a(x), p(x) and q(x) are given functions of the transverse co-ordinate
x, x [h/2, h/2]. In the next sections it will be shown that a(x) =

Figure 7.1: Displacements described by the internal degrees of freedom.

2 sin(x/h), p(x) = 2 cos(2x/h) and q(x) = 2 /24 sin(x/h). Thus,


functions h( + ) describe the amplitudes of the sinusoidal shear of the
transverse fibers, measures their sinusoidal extension, while v correspond
to the amplitudes of the complicated motion of the transverse fibres, when
the particles on the face surfaces move in one direction and those of the
middle surface in the opposite one (see Figure 7.1).
From the phenomenological point of view the density of the 2-D kinetic
energy appears to be natural. The density of the 2-D strain energy has
some special features, namely, i) the cross terms between A and , A
and v; , and v are absent, ii) the higher derivatives of the unknown

254

CHAPTER 7. ELASTIC SHELLS

functions do not enter . The latter is distinctive in comparison with the


classical shell theory; it means that the functional (7.1)-(7.3) is defined on
the space of all admissible functions u , u, , and v which need be only
continuously differentiable.
The derivation of (7.2) and (7.3) as well as the determination of the
constants (7.5) will be given in the next two sections. It will be shown that
all the three requirements posed to the theory of high-frequency vibrations
for shells are satisfied.
2-D equations of motion. Now let us derive the 2-D equations of highfrequency vibrations for the elastic shell according to the variational principle
(7.1). Assuming the values of the functions u , u, , and v to be specified
at t = t0 , t1 and varying the action functional with and from (7.2) and
(7.3), we have
Z t1 Z
+ v v ) n A
J =
[h(u
u + u u + h2 +
S

t0

m q ( + u, + b u ) h(s2 ; , + s3 v;
v;

+ 2v (;) v; + r1 h1 A + r2 h1 v;
+ r2 h1 v;
+ r3 h1 v ,

+ r3 h1 ; v + f2 h2 + f3 h2 v v )] da dt,
where the following quantities are introduced
n =

= h(s1 A a + 2A + r1 h1 a ),
A
h3

= ( a + ),
m =

(7.6)

q = hf1 ( + u, + b u ).
Transforming J by the integration by parts and using the conditions of
vanishing variations at t0 and t1 , one can show that
Z t1 Z

u + q;
+ t b )u
J =
{(h
u + t
; q b )u + (h
t0

+ (h3 q m
; ) + [h

+ h(s2 2 r1 h1 A r23 h1 v;
f2 h2 )]
2

+ 2(v(;) ); + r23 h1 , f3
h v )]v } da dt
+ [h
v + h(s3 v;
Z t1 Z
+
[t u q u + m h(s2 ; + r3 h1 v )
t0

S

h(s3 v;
a + 2v (;) + r2 h1 a ) v ] ds dt, (7.7)

7.1. TWO-DIMENSIONAL EQUATIONS

255

where

1
t = n + (b m b m ),
2
the constant r23 = r2 r3 and 2 is the 2-D Laplace operator. Due to
the arbitrariness of the variations of u , u, , and v inside S (t0 , t1 ) the
variational equation J = 0 yields

h
u = t
; q b ,

h
u = q;
+ t b ,
h3 = q m ,

(7.8)

f2 h2 ),
h = h(s2 2 r1 h1 A r23 h1 v;
2

+ 2(v(;) ); + r23 h1 , f3
h v ],
h
v = h[s3 v;

These are the 2-D equations of high-frequency vibrations for elastic shells.
Boundary-value problems. Let us consider the variant of the boundary
conditions for the system of equations (7.8) that correspond to the free edge.
In a manner similar to the classical shell theory we assume that the variations
of all the unknown functions are arbitrary at S. Then from (7.1) and (7.7)
the natural boundary conditions follow
t = 0, q = 0, m = 0,
(s2 ; + r3 h1 v ) = 0,

(7.9)


(s3 v;
a + 2v (;) + r2 h1 a ) = 0.

The other interesting variant of the boundary conditions for (7.8) can be
obtained if we assume, for instance, that u, , u2 , 2 and v2 vanish at S,
but the remaining functions may be varied arbitrarily there. In this case the
boundary conditions read
u = 0, = 0,
u2 = 0, 2 = 0, v2 = 0,
11
t = 0, m11 = 0, v 1;1 = 0,

(7.10)

where the index 2 denotes the tangential direction. It will be seen later which
physical situation these boundary conditions correspond to.
If the tractions act on the face surfaces, then the variational principle
(7.1) should be modified by adding the term associated with the work done
by the surface tractions:
Z t1 Z
J =
( + A) da dt = 0,
(7.11)
t0

256

CHAPTER 7. ELASTIC SHELLS

where
A = {t }u + {t}u + a1 h[t ] + a2 h[t ]u, + a3 [t] a4 {t }v ,

(7.12)
a1 = 2 /24, a2 = 1/2 + 2 /24, a3 = a4 = 2.
In (7.12) {t} = t+ +t and [t] = t+ t . From (7.11) and (7.12) the following
equations can be derived

h
u = t
; q b + {t },

h
u = q;
+ t b + {t} a2 h[t; ],
h3 = q m + a1 h[t ],
;

h = h(s2 2 r1 h1 A r23 h1 v;
f2 h2 ) + a3 [t],
2

h
v = h[s3 v;
+ 2(v(;) ); + r23 h1 , f3
h v ] a4 [t ].

The constitutive equations (7.6) and the boundary conditions remain the
same.
To complete the formulation of the boundary-value problems we specify
also the initial conditions at t = t0
u |t0 = u0 , u|t0 = u0 , |t0 = 0 , |t0 = 0 , v |t0 = v0 ,
t = 1 , v |t = v 1 .
u |t0 = u1 , u|
t0 = u1 , |t0 = 1 , |
0
0

(7.13)

If the system of equations (7.6) and (7.8) is of the hyperbolic type with
respect to the unknown functions u , u, , and v, then they compose together with the boundary conditions (7.9) and the initial conditions (7.13)
the well-posed initial boundary-value problem. We shall return to this question in the next sections.
Problems
1. Show that the equations of motion and the boundary conditions for a
plate (with b = 0) break up into those of longitudinal and flexural
vibrations. Find the corresponding equations and boundary conditions
in terms of u , , v and u, .
2. Derive the balance equation of energy for a shell within the theory of
high-frequency vibrations. Using this equation, prove the uniqueness
of solutions of the boundary-value problem.
3. Formulate the variational principle for the eigenvalue problem and
prove the orthogonality of the eigenfunctions with respect to the energy.

7.2. LONG-WAVE ASYMPTOTIC ANALYSIS

7.2

257

Long-wave asymptotic analysis

Thickness vibrations near the cut-off frequencies. In the previous


section the 2-D theory of high-frequency vibrations for elastic shells has been
proposed. According to the requirements formulated at the very beginning
of that section, this theory should describe correctly the low-frequency and
some of the first high-frequency branches of vibrations of the shell in the
long-wave range. This means that the degrees of freedom introduced should
correspond to the eigenfunctions of the classical and thickness branches of
vibrations in the long-wave range. The long-wave state in shells is defined
as the state whose smallest wavelength l of the deformation patterns in the
longitudinal directions is considerably greater than the shell thickness h. The
possible types of long-wave vibrations can be classified roughly as follows. Let
the face surfaces of the shell be traction-free. Since l  h, the derivatives of
the displacements with respect to the longitudinal coordinates x ( = 1, 2)
can be neglected in the three-dimensional equations of motion and in the
traction free boundary conditions as small compared with the derivatives
with respect to the transverse coordinates x [h/2, h/2]. Then the system
of equations and boundary conditions breaks up into the following systems:
w =

h
2 w
, |x| ,
2
x
2

w
h
= 0, x = ,
x
2

(7.14)

and

2w
h
w
h
, |x| ,
= 0, x = .
(7.15)
2
x
2
x
2
Here w and w are the components of the displacements referred to the basis
{t , n}. The complete set of particular solutions of (7.14) and (7.15) follows

w = v 2 cos , w = 0, = 2n,
(F (n)),
(7.16)

w = 0, w = 2 sin , = (2n + 1), (Fk (n)),


(7.17)

w = 2 sin , w = 0, = (2n + 1), (L (n)),


(7.18)

w = 0, w = v 2 cos , = 2n,
(Lk (n)),
(7.19)
w = ( + 2)

with = x/h. The quantities and run through a countable set of values;
however, no indices are attached to
and in order to avoid complicated
notation. The normalization factor 2 is chosen so as to simplify the twodimensional kinetic energy. It is understood that the functions v, , , and
v correspond to each value of or ; these functions are also not numbered.
Furthermore, they depend harmonically on t with frequency determined
by appropriate values of or from the formulae
cd
cs
=
or =
.
(7.20)
h
h

258

CHAPTER 7. ELASTIC SHELLS

Here cd and cs are the velocities of dilatational and shear waves, respectively.
The notation for series of different solutions is indicated in parentheses in
(7.16)-(7.19), where the symbols F and L stand for flexural (antisymmetric)
and longitudinal (symmetric) vibrations, and the subscripts and k denote
the thickness-stretch and thickness-shear branches, respectively.1
For functions v, , , v independent of x , each of the solutions given
above represents an exact solution of the equations of 3-D elasticity for an
infinite plate and corresponds to the synchronized vibrations of transverse
fibres along the plate (with the zero longitudinal wave number). The frequencies (7.20) will be called cut-off frequencies. For vibrations whose amplitude
and frequency vary slowly in the longitudinal directions of the plates and
shells, the equations (7.14)-(7.15) can be regarded as the zero approximations. The solutions (7.16)-(7.19) can be considered as the principal terms in
a certain asymptotic expansion in which v, , , v are functions of x and
t, where
v 2 v, 2 , 2 , v 2 v .
(7.21)
The values of in these estimations are taken for the same branch as the
corresponding function, with the exception of F (0) and Lk (0), for which it
is assumed that v cs v/l, v cs v /l, where l is the smallest wavelength
of the deformation patterns. The branches F (0) and Lk (0) correspond to
the low frequency vibrations when h/cs  1. The independence of the
displacements of these branches from x in the zero approximation is a part
of the Kirchhoff-Love hypotheses. All the remaining branches correspond to
vibrations with frequency cs /h. The propagation time for a perturbation
over the thickness for these branches of vibrations is of the same order as the
period of vibrations; therefore it is impossible to assume the displacements
polynomials in x even in a zero approximation. Since as h 0, the
corresponding vibrations are naturally called high-frequency (or thickness)
vibrations.
Taking for example n = 1, cs = 2500m/s (steel, e.g.), h = 3mm, we
have 1 ' 4.105 Hz for the branch Fk (0), i.e, 1 is in the ultrasonic domain.
Vibrations of elastic bodies at such a high frequency may be important in
problems of impact or in problems of vibrations caused by piezoelectric resonators. Note that for layered shells of sandwich type with a significant drop
in the elastic moduli, 1 is considerably smaller and can even be in the audio
frequency domain [51]. The high-frequency branches have the displacement
distributions which oscillate all the more rapidly over x as n grows. The
distribution of the branch n has 2n or 2n + 1 nodal points over the thickness.
Let us mention that the wavelength of the high-frequency branches in the
1

This terminology was introduced by Mindlin [41].

7.2. LONG-WAVE ASYMPTOTIC ANALYSIS

259

transverse direction of the three-dimensional shell is smaller than h, but this


fact is not an obstacle for the application of the asymptotic analysis, which is
based on the smallness of h/l with l being the wavelength in the longitudinal
directions.
3-D variational formulation. To formulate the problem of free vibrations of the shell we refer its undeformed state to the curvilinear coordinates
{x , x}
z i = ri (x ) + xni (x ).
In this coordinate system the co- and contravariant components of the metric
tensors are given by the formulae (3.33) and (3.36). However, for the purpose
of the asymptotic analysis it is enough to approximate g by
g = a + 2b x + 3c x2 + O(

h3
)a .
R3

According to Hamiltons principle the displacement field of the free vibrations of the shell is a stationary point of the action functional
Z

t1

Z Z

h/2

(T W ) dx da dt,

I=
t0

(7.22)

h/2

where = 1 2Hx + Kx2 , with the Lagrangian T W given by


1
1
T W = (w 2 + a w w ) [ (g  + 33 )2
2
2

2
+ g g   + (33 ) + 2g 3 3 ].
As before, the variations of w , w should vanish at t0 , t1 . Concerning the
boundary conditions at the shell edge, we first consider the clamped edge,
for which
w = 0,

w = 0 at S (h/2, h/2).

Variational-asymptotic procedure. It turns out that for high-frequency


long-wave vibrations the variational problem of finding the extremals of
the functional (7.22) can also be studied with the help of the variationalasymptotic method. We assume the smallness of the parameters h/R and
h/l everywhere in S, but we do not assume that the frequency of vibrations is
small. Making use of the change of variable = x/h and discarding formally

260

CHAPTER 7. ELASTIC SHELLS

all small terms in (7.22) we arrive at the zero approximation functional


Z t1 Z Z 1/2
1
1
[ w 2 + a w w
I0 = h
2
S 1/2 2
t0
1
1
( + 2) 2 (w| )2 2 a w| w| ] d da dt.
2h
2h
One can see that, in contrast to the case of low-frequency vibrations, the
kinetic energy must already be retained in the first step of the variationalasymptotic procedure since the frequencies are no longer small. The Euler
equations of this functional yield four series of free vibrations (7.16)-(7.19).
It is convenient to introduce further in the series F and L the number
according to
r
r

1 2
cs
=
=
.
(7.23)
= e, e =
cd
+ 2
2 2
Similarly, we also introduce in the series Fk and Lk the number by the
same formula (7.23).
Now we find the next refinement for the displacements of branches of the
series F . Considering v belonging to the branch F (n) a given function of
x and t satisfying the constraint
v = 0 at S,

(7.24)

we seek w for that branch. Keeping the principal terms depending on w


and the principal cross terms in (7.22), we obtain the functional
Z t1 Z Z 1/2
I1 = h
1 d da dt,
(7.25)
t0

1/2

with

1
1 = 2 w w w v, h1 2 sin
2

1 1
a (h w| + v, 2 cos )(h1 w| + v, 2 cos ).
2
Integration by parts was performed and terms that go to the boundary vanish
due to (7.24). Thus, the functional does not depend on the derivatives of
w with respect to x and t, and these enter it as parameters. Let us find
the stationary point of this functional. Varying (7.25) with respect to w we
obtain the boundary-value problem

+
w| + 2 w =
hv, 2 sin , || 1/2,

w| + hv, 2 cos = 0, = 1/2.

7.2. LONG-WAVE ASYMPTOTIC ANALYSIS

261

It yields the following tangential displacements


w = v,

h
2(1)n e sin()
2(sin
).

cos(/2)

(7.26)

As expected, the tangential displacements turn out to be much smaller than


the normal displacement in the long-wave range and are of the order hu/l.
Let us seek the correction to w

w = v 2 cos + w? .
Here w is considered fixed and defined by (7.26). By redefining v if needed,
the following constraint can be imposed on w?
Z

1/2

w? cos d = 0.

1/2

This corresponds to the assumption that v = hw 2 cos i, where h.i denotes the integration over from -1/2 to 1/2. After discarding small terms
containing w? and small cross terms as compared with those remaining, the
functional (7.22) takes the form (7.25) with a Lagrangian given by

1
1
1 = 2 (w? )2 2 2Hhv 2 cos w? 2 ( + 2)(w|? )2
2
2h

2H
2H ?
( + 2)
2Hv 2 sin w|? +
w,? v 2 cos
w v 2 sin .

h
h
h
Its stationary point has the form

1 4e2
w? = Hhv 2( cos +
sin ).

Summing up, we have the following distribution of the displacements over


the thickness in the series F (within the first approximation)
F :

1 4e2
w = v 2 cos + Hhv 2( cos +
sin ),

h
2(1)n e sin()
w = v,
2(sin
).

cos(/2)

(7.27)

Now let us consider the series Lk . Fixing v , we seek w in the second


step for the branch Lk (n) (in the first step w = 0). Keeping the principal

262

CHAPTER 7. ELASTIC SHELLS

terms of w and the principal cross terms between w and v , we obtain the
functional (7.25) with

1
1

w| 2 cos
1 = 2 w2 ( + 2)h2 w|2 h1 v;
2
2
1
h v; w 2 sin .
The last term of this equation was also obtained after the integration by
parts taking into account that v vanish at the boundary. The stationary
point is easily found to be

w = v;

h
2(1)n e sin()
).
2( sin +

cos(/2)

We seek the refined term to w in the form

w = v 2 cos + w? ,
where w? satisfy the following constraints
hw? cos i = 0.
Again, after discarding small terms containing w? and small cross terms as
compared with the remaining terms, the functional (7.22) takes the form
(7.25) with a Lagrangian given by

1
?
1 = 2 (w? )2 2 2Hhv 2 cos w? 2 (w|
)2
2
2h

?
?
v 2 cos + b w? v 2 sin .
b w|
2Hv 2 sin w|
h
h
h
Its stationary point is found to be
b

w? = h 2(Hv cos v sin ),

where b = b + H . Thus, for the series Lk we have


Lk :

w = v 2 cos + h 2(Hv cos v sin ),

2(1) e sin()
h
w = v;
2( sin +
).

cos(/2)

(7.28)

7.2. LONG-WAVE ASYMPTOTIC ANALYSIS

263

Analogously, formulae are obtained for the displacements in the two remaining series
Fk :

L :

w = 2 sin + h 2(H sin + cos ),

2(1) e cos()
h
w = ;
2(cos
),

sin(/2)

(7.29)

1 4e2
w = 2 sin + Hh 2( sin
cos ),

2(1)n e cos()
h
2( cos +
).
w = ,

sin(/2)

(7.30)

The distinguishing feature of shells as compared with plates [6] is that the
correction terms in the displacements are of the order h/R compared with the
principal term, while they are of the order (h/l)2 in plates. By continuing the
iteration process, the next corrections to w and w can be found. They are
not presented here since they yield no contribution to the average Lagrangian
of the first approximation.
Average Lagrangian of thickness branches of vibrations. Let the displacements w, w be expressed by the infinite series of branches given above,
where v, , , v are arbitrary functions of x and t. After substituting these
series into the action functional (7.22) and integrating over the thickness we
neglect those small terms of the order h/R, h/l compared with 1. It turns out
that the thickness branches are orthogonal relative to the action functional
in the long-wave range, provided the shell edge is clamped [6, 7]. Therefore
the average functional has the form
Z

t1

J =h
t0

da dt,

(7.31)

decomposes into the series of average Lawhere the average Lagrangian


grangians of low frequency and thickness branches.2 For the series F we
get
F :

= v 2 + l2 (h/)2 a v , v , + l4 (h/)2 v 2
2
( + 2)((/h)2 v 2 + l1 a v, v, + l3 v 2 ),

The average Lagrangian of the low frequency branches is given in Section 3.2.

264

CHAPTER 7. ELASTIC SHELLS

where
2e2 tan(/2) 5 3e2
2e2
),
+
/2
1 e2
cos2 (/2)
3 e2 4e2 tan(/2)
4e2
l2 = 1
+
,
1 e2
/2
cos2 (/2)
3 2
8e2 ) + 4H 2 (1 5e2 + 4e4 ),
l3 = (3H 2 K)( +
2 12
1 2
2
l4 = (3H K)( + ) + 2H 2 (1 6e2 + 8e4 ).
2 12

l1 = 2(1

Within the first-order approximation one can further simplify this expression. Indeed, transforming the terms l2 (h/)2 a v , v , and l4 (h/)2 v 2 by
integration over t by parts and neglecting terms3 that go to the boundaries
t = t0 , t1 and do not affect the Euler equations inside (t0 , t1 ) we obtain instead
of them
l2 (h/)2 a v, v, l4 (h/)2 vv.
Because these terms are small correction terms in the average Lagrangian,
and because near the cut-up frequency equation (7.21) holds true, one can
replace v by 2 v. Now the average Lagrangians for the series F become
F :

= v 2 [(h2 2 + k2 )v 2 + k1 a v, v, ],
2

with
16 tan(/2)
1

,
e2

1
1
k2 = H 2 ( 2 16) + K( 2 8).
e
e
k1 =

Turning to the average Lagrangian of the series Fk we have


Fk :

2
= (a + (h/)2 l4 ) + l2 (h/)2 ( ;
)
2

2
) ],
[((/h)2 a + l3 ) + 2(;) (;) + l1 a (;

One already observes here some kind of short-wave extrapolation.

(7.32)

7.2. LONG-WAVE ASYMPTOTIC ANALYSIS

265

where
4e2
1 3e2 sin
),
(1
+
1 e2
sin2 (/2)
4e2
1 + e2 sin
),
l2 = 1 +
(1

1 e2
sin2 (/2)
2 1
l3 = (K 3H 2 )a ( ) + 3(3b H Ka ),
12 2
2 1
l4 = (K 3H 2 )a ( + ) + 3b H Ka .
12 2
l1 =

With the same deliberations one can replace the terms (h/)2 l4 and
2
) in the long-wave range by
l2 (h/)2 ( ;
l4

2
and l2 (;
),

respectively. Then the average Lagrangian reduces to


Fk :

= a [(h2 2 a + k2 )
2
2
+ 2(;) (;) + k1 (;
) ],

(7.33)

with
16e2 cot(/2)
,

= (3H 2 K)a + 6Hb 2Ka ,

k1 = 1 +
k2

In a similar way we obtain the following average Lagrangians for the


remaining two series
L :

= 2 [(h2 2 + k2 ) 2 + k1 a , , ],
2

with
k1 =

(7.34)

1
16 cot(/2)
+
,
2
e

and
Lk :

= a v v [(h2 2 a + k2 )v v
2
2
+ 2v(;) v (;) + k1 (v;
) ],

with
k1 = 1

16e2 tan(/2)
.

(7.35)

266

CHAPTER 7. ELASTIC SHELLS

The coefficients k2 in the series L and the tensors k2 in the series Lk are not
written down here since they are identical in form with those in the series F
and Fk . Not only the principal terms containing the factor 1 in the kinetic
energy and the factor h2 in the strain energy but also terms of the next
order of smallness must be retained in the average Lagrangians (7.32)-(7.35),
due to the fact that, at the cut-off frequencies, the sum of the principal terms
turns out to be small.
from (7.32)-(7.35) we arrive
By varying the action functional (7.31) with
at the following equations of thickness vibrations
F :

v = [(h2 2 + k2 )v + k1 v],

(7.36)

and
Fk :

)
= [(h2 2 + k2

+k1 ;
+ 2((;) ); ].

(7.37)

The equations for the series L and Lk can be obtained from (7.36) and
(7.37) by making the respective substitutions: v , and v . These
equations coincide with those of [21] derived by the asymptotic method of
Goldenveizer.

Figure 7.2: Graph of k1 as the function of .


It is interesting to note here that the type of equations (7.36)-(7.37) depends on the coefficients k1 . For instance, equation (7.36) is of the hyperbolic
type if k1 > 0. Figure 7.2 shows the graphs of k1 as the functions of the Poisson ratio for the branch L (0). One can see that for the range 0 < < 1/3
this coefficient is negative, and the equation of vibration is of the elliptic
type. For = 1/3 the coefficient k1 = , and (7.36) becomes ill-defined.

7.3. SHORT-WAVE EXTRAPOLATION

267

This shows clearly that the correct asymptotic analysis does not always lead
to well-posed boundary value problems.
We shall see in the next section how this pathological feature of the
equations of thickness vibration could be removed by extrapolating them to
short waves taking into account the cross terms between branches.
Problems
1. Follow the variational-asymptotic procedure to derive the distributions
of displacements (7.29) and (7.30) for the series Fk and L , respectively.

2. Derive the asymptotic distributions of displacements for plates from


the exact solution (3.74) near the cut-off frequencies.
3. Calculate the coefficients of the average Lagrangians for the series
Fk , L and Lk .
4. Apply equations (7.36) to determine eigenfrequencies of the axisymmetric radial vibrations of closed spherical shells.

7.3

Short-wave extrapolation

Derivation of the theory of high-frequency vibrations. Let us consider free vibrations of the shell, now with arbitrary boundary conditions
at its edge (clamped or free edge). We assume that these vibrations can
be regarded with sufficient accuracy as the superposition of the branches
F (0), Fk (0), Lk (0), L (0) and Lk (1). The branches F (0) and Lk (0), in the
long-wave range, correspond to low-frequency vibrations, the other ones to
thickness vibrations with the lowest frequencies. Such a choice is based on
the following reasoning. First, these branches possess the lowest cut-off frequencies,4 and, therefore, the most essential part of the vibrational energy
is concentrated in them. Second, the necessity of including also Lk (1) into
the theory is dictated by its strong interaction with the branch L (0) [41].
The dynamic equations contain eight unknown functions of the longitudinal
v (the symbols without the bar are recoordinates and time: u, u , , ,
served for the functions in the final equations). Despite the fact that the
theory involves more unknown functions than in the classical shell theory,
it should be regarded as a first approximation theory describing asymptotically exactly the vibrations of the shell in the range of long waves and high
frequencies ( 2cs /h).
4

This is true only for (0, 7/16), and we shall restrict ourselves to this condition.

268

CHAPTER 7. ELASTIC SHELLS

Thus, we represent the displacements of the shell in the form

w = u h A + h2 l() + a()
+ h
v;
g() + h;
q(),
w = u h + h3 m() + h3 n() + p ()
;

+
v d ()

+ h, f ().

(7.38)

In the right-hand side of equations (7.38) the first three terms of the first
equation and the first four terms of the second one describe the low-frequency
branches (cf. (3.51)).5 The 2-D measures of extension and bending are
expressed through u , u as follows
A = u(;) b u,
= (;) + b( $
) ,
1
u, u, ).
= u, + b u , $
= (
2
The functions l(), m(), n() have the form
1
3
1
l() = ( 2 1/12), m() = ( 3 ),
2
3
4
1 3 5
n() = ( ),
6
4
where = /(1 ). The functions a(), d (), f (), g() are given by the
formulae (7.28) and (7.30), which, in our case, read

1 4e2
cos ),
2 sin + Hh 2( cos

d () = 2 cos 2 + h 2(H cos 2 sin 2),


2

2
2e cos(/e)
f () =
( cos +
),

sin(/2e)

2
2e sin(2e)
g() =
( sin 2
),
2
cos(e)
a() =

(7.39)

For the functions p () and q() it is convenient to use another normalization


5

The third and fourth terms in the right-hand side of (7.38)2 should be included for
the consequent account of all principal cross terms in the average Lagrangian. They
can be obtained as the correction terms in the third step of the variational-asymptotic
procedure [6].

7.3. SHORT-WAVE EXTRAPOLATION

269

factor. We therefore represent them in the form


p () = c sin + hc(H sin +
q() =

sin ),

c
2e cos(e)
(cos
),

sin(e/2)

with c at the moment an undefined normalization constant that will be chosen


later to simplify the subsequent changes of unknown functions.
Knowing the displacement field, we now calculate the main terms of the
strain field
 = A h
+ (;) p + v(;) b + h; f,
23 = h2 m| + h2 (n| + l) + h1 p
;

+ h v d| + , (f| + a) +
|
33 = A + h + h1 a
1

h
v;
g,

v;
g|

(7.40)
+

;
q| ,

where b() = 2 cos(2) and p() = c sin(). The vertical bar preceding
denotes the derivative with respect to . In the expression for 3 in

2
(7.40) we omit the term h;
q, which gives a contribution (;
) to the
average Lagrangian. This is due to an additional analysis, which shows that
a hyperbolic short-wave extrapolation describing exactly the curvature of the
dispersion curve near the cut-up frequency of the branch Fk (0) does not exist.
We substitute the formulae (7.38) into the action functional (7.22) and
integrate over the thickness. Discarding small terms in the asymptotic sense
and using the results of Chapter 3 and the previous section, after long but
otherwise standard calculations one obtains the average functional (7.31)
with
2
= 1 (u 2 + c 2 + u 2 + 2c1 hu ; + 2 + 2c2 h A + 2c3 h v + v 2

2
2
1
v
+ 2c4 hv ; ) [22 h2 2 + 32 h2 v2 + 2d1 h1 A + 2d2 h1
;
2
2
2
) + 2
v (;) v(;)
+ k3 (
v;
+ 2d3 h1 v , + 2(A )2 + 2A A + k2 ,

c2
h2

((
)2 + ) + 2hd4 ;
+ 2hd5 (;) + h2 12 2
6
2



2
2

+ 2hd6 ; + 2hd7 ; + (c /2)s4 + s5 + s4 v v ]. (7.41)


+

In this Lagrangian the coefficients 1 , 2 , 3 and k2 , k3 , s


4 , s5 are given by

270

CHAPTER 7. ELASTIC SHELLS

the following formulae

2 = , 3 = 2,
e

1
8e
16e
k2 = 2 +
cot
, k3 = 1
tan(e),
e

2e

s
+ 6Hb 2Ka ,
4 = (3H K)a
s5 = H 2 (1/e2 16) + K(1/e2 8).
1 = ,

(7.42)

The remaining coefficients are expressed through l, m, n, p, q, a, b, f, g by


c1 = hf i, c2 = hai, c3 = hagi, c4 = hbf i,
d1 = ha| e2 a| i, d2 = ha| b e2 a| g| i,
(7.43)
2
d3 = hb| (f| + a)i, d4 = hp + l| p q| + e l| q| i,
d5 = 2hpi, d6 = hp| (n| + l)i, d7 = hp| m| i.
The direct calculations show that



2 2
1
2e2
4
c1 = c2 = 2 , c3 = c4 = 2 + 2
,

3 e 1/4
d1 = 0, d2 = d3 = 22 c3 + r2 = 32 c3 + r3 ,
4
16
4c
4c
r2 = 2 , r3 = , d4 = d6 = 2 , d5 = d7 = 2 .
3e
3

(7.44)

Note that the identities c1 = c2 , c3 = c4 , d1 = 0, d2 = d3 , d4 = d6 , d5 = d7 can


be proved without calculating the integrals in (7.43). This expresses the fact
that all the cross-terms in (7.41) form divergence terms that do not affect the
equations of vibrations in the long-wave range and lead to the orthogonality
between branches when the edge is clamped. In the calculus of variations such
terms are called null Lagrangian that is, a Lagrangian whose Euler operator
vanishes identically. Consequently, a method of short-wave extrapolation is
possible in which all the cross terms in (7.41) are discarded, which would lead
to the uncoupled equations (7.36) and (7.37). However, additional analysis
shows that such extrapolation leads to a qualitatively false description of the
dispersion curves and integral characteristics of shells and plates in the shortwave range (see Section 8.4). Here we shall not neglect the cross-terms, but
merely look for a change of unknown functions, which enables one to get rid
of terms containing the higher order derivatives by making them negligibly
small in the long-wave range. Using (7.44) we transform the Lagrangian

7.3. SHORT-WAVE EXTRAPOLATION

271

(7.41) to
2
= 1 [u 2 + c 2 + (u + c1 h , )2 + ( + c1 hA + c3 hv )2

;
2
2
1
v )2
+ (v + c3 h , )2 ] [22 h2 ( + c1 hA + c3 h
2
v + 2r3 h1 v ,
v + c3 h, )2 + 2r1 h1 A + 2r2 h1
+ 32 h2 (
;
2 2
2

2 2
2

+ (2 2 c1 )(A ) + 2A A + (k2 + c1 2 + c23 22 c23 32 ),

2
2
+ (k3 + c23 32 c23 22 )(
v;
) + 2
v (;) v(;) + (h
)
+ 6d5 ;
6
1
c2
+ (h
+ 6d5 (;) )(h
+ 6d5 (;) ) + h2 12 (
6
2
d5 3
d5 3 2


2
+ 2 h ; + 2 h ; ) + (c2 /2)s
v ], (7.45)
4 + s5 + s4 v
c
c

where r1 = 12 c1 = 2 2/e2 . While transforming (7.41) to (7.45) terms of

2
the type h2 (A )2 , A v;
, (;
) and h4 (
; )2 are neglected as small compared
with the remaining terms in the long-wave range. Besides, the small correc2
2
tion terms of the type h2 ,
and h2 (v ;
) are replaced by the asymptotically
2 2
2 2
equivalent terms 2 , and 3 (v; ) , as in the previous section.
In order to search for a short-wave extrapolation which does not contain
second and higher derivatives in the Lagrangian let us choose6 the constant
c = 2 /24 and make the following changes of unknown functions

u = u,
u = u + c1 h, ,
24
1 24
= + h1 ( ( 3 )2 h3 ; ( 3 )2 h3 ; ),
3
3

= + c1 hA + c3 h
v; ,
v = v + c3 h, .

(7.46)

The sense of these changes of unknown functions is to make all terms containing second derivatives of the new functions negligibly small in the long-wave
range. The Lagrangian (7.45) then becomes
= 1 (u 2 + h2 2 + u 2 + 2 + v 2 ) 1 [ 2 h2 2 + 2 h2 v 2

2
3

2
2

+ 2r1 h1 ( c1 hA c3 hv;
)(A c3 h2 ) + 2r2 h1 ( c1 hA
6

It is easy to show that the Lagrangian does not depend on this special choice.

272

CHAPTER 7. ELASTIC SHELLS

c3 hv;
)(v;
c3 h2 ) + 2r3 h1 (v c3 ; )(, c1 hA, c3 hv;
)
2
+ (2 22 c21 )(A )2 + 2A A + (k2 + c21 22 + c23 22 c23 32 ),

h2
(( )2 + )
6
+ f1 ( + )( + ) + s5 2 + s
4 v v ], (7.47)

2
+ (k3 + c23 32 c23 22 )(v;
) + 2v (;) v(;) +

where = c2 /2 = ( 2 /24)2 /2, f1 = ( 2 a + h2 s


4 ). The new measures
A and are expressed through u, u and according to (7.4). While
changing to the new unknown functions terms of the type h2 2 , h2
are neglected as small in the long-wave range. Terms of the type (A )2 ,
2 2
2
(
v;
) , , are replaced by the asymptotically equivalent terms (A )2 , (v;
),
2
, . Keeping in (7.47) the principal terms and replacing terms of the type
2
2

) , respectively (with the help of the


and (v;
by ,
2 and v v;
integration by parts or, equivalently, by adding a null Lagrangian), we arrive
finally at the formulae (7.2) and (7.3), where
s1 = 2 22 c21 2r1 c1 ,
s2 = k2 + c21 22 + c23 22 c23 32 + 2r1 c1 + 2r2 c3 2r3 c3 ,
s3 = k3 + c23 32 c23 22 2r2 c3 + 2r3 c3 .
Making use of (7.42) one can obtain all the formulae (7.5) for the constants
of the theory.
Thus, the formulae of the 2-D theory of high-frequency vibrations is justified by the short-wave extrapolation procedure. The latter consists of the
operations of adding or removing i) terms that are small in the long-wave
range, or ii) null Lagrangians. One can see that such a procedure leaves
the theory asymptotically exact for long waves, but may improve essentially
the behaviour of the differential operator for short waves. In our case, the
obtained by the short-wave extrapolation is a quadratic form
Lagrangian
with respect to u , u, , , v and their first derivatives, and it turns out
that the corresponding Euler equations are hyperbolic.
Consider the case when the traction acts on the face surfaces. To modify
the 2-D theory we must calculate the work A done by the surface traction ti .
Keeping only the principal terms, we obtain for A the following expression
h
A = {t }
u + {t}
u [t ] + [t ] q(1/2)
2

+[t]a(1/2) + {t }
v p(1/2).
From the formulae (7.39) one can see that

q(1/2) = 2 /24, a(1/2) = 2,

p(1/2) = 2.

(7.48)

7.3. SHORT-WAVE EXTRAPOLATION

273

Changing to the new unknown functions and keeping the main terms in (7.48)
we get for A the formula (7.12).
Hyperbolicity of the 2-D equations. Let us show that the 2-D theory of
high-frequency vibrations for shells obtained by the short-wave extrapolation
satisfies all the requirements posed at the beginning of Section 7.1. First, the
linearity of the 2-D equations is obvious due to the quadratic Lagrangian.
Second, the asymptotical exactness follows from the variational-asymptotic
procedure. What remains to prove is the hyperbolicity of the 2-D equations.
The type of the equations is characterized by terms containing the highest
derivatives [44]. Concerning the Euler equations of the functional (7.1) we
= depends only on
shall use the following result: if the Lagrangian
the unknown functions and their first derivatives with respect to x and t,
and if both the quadratic forms containing the time and spatial derivatives in
and are positive definite, then the system of equations of the functional
(7.1) is hyperbolic. It is obvious from (7.3) that is the strictly positive
definite quadratic form of u,
u , , and v . We now select in the
quadratic form that contains only the first derivatives with respect to x ,
which is given by
h2

h[s1 (u, )2 + 2u(,) u(,) + (( )2 + )


2
6
2
2
+ s2 ,
+ s3 (v,
) + 2v (,) v(,) + 12 u2, ],

1 =

where = (,) + b( $) . It can be seen that 1 is positive definite if


s1 + 1 > 0,

+ 1 > 0,

s2 > 0,

s3 + 1 > 0.

The first two inequalities always hold true for [0, 1/2] ( is Poissons
ratio). Consider now the last two inequalities. According to (7.5) we have
s2 =

16e cot(/2e)
16(1 + 2e2 )2
8 2
1
+
+

,
e2

9 2 (4e2 1)e2 2 e2
8e tan(e)
16(1 + 2e2 )2
s3 + 1 =
2 2
.

9 (4e 1)e2

Thesepcoefficients depend only on Poissons ratio , since = /(1 ) and


e = (1 2)/(2 2). For real elastic materials [0, 1/2]. One can
prove that s3 + 1 > 0 for [0, 1/2], and s2 > 0 for [0, 0.44). Figure 7.3
depicts the graphs of s2 and s3 + 1 as functions of in the interval (0,0.44),
where one can see that both are positive in this interval. Thus, we have
shown that for [0, 0.44) the quadratic form 1 is positive definite and

274

CHAPTER 7. ELASTIC SHELLS

Figure 7.3: Graphs of s2 and s3 + 1 as the functions of .


the Euler equations (7.8) are hyperbolic. As a consequence, one can also
prove the well-posedness of the corresponding initial boundary-value problems. For (0.44, 1/2) the coefficient s2 is a quickly oscillating function
of , which changes its sign infinitely many times. One can explain this by
the fact that, as increases, the cut-off frequency 2 = /e of the branch
L (0) also increases and crosses the regions of strong interaction with other
high-frequency branches (Lk (2), Lk (3), and so on). Due to this reason the
short-wave extrapolation proposed above becomes unsatisfactory. In contemplating to include other branches, which interact strongly with L (0),
one should remember that each additional branch increases the order of the
differential equations by two. Due to the increasing complexity, it would
appear preferable to use the finite element method within 3-D elasticity to
determine the eigenfrequencies in such situations.
Reduction to the theory of low-frequency vibrations. It turns out
that in the low-frequency range h/cs  1 the theory of high-frequency
vibrations (7.1)-(7.3) is asymptotically equivalent to the classical theory of
low-frequency vibrations. This statement can be proved by the variationalasymptotic method. Indeed, since the functional (7.1)-(7.3) contains the
small parameter h, one can study its asymptotic behaviour as h 0 and
h/cs 0. As h/cs 0, in the first step of the variational-asymptotic
procedure the kinetic energy can be neglected. Keeping the main terms in
the strain energy, we reduce the functional to
h
J0 =
2

t1

+ 2r3 h1 v ,
[2r1 h1 A + 2r2 h1 v;

f1 (

t0

+ )( + ) + f2 h2 2 + f3 h2 v v ] da dt. (7.49)

It is easy to see that (7.49) determines the following asymptotic formulae as

7.3. SHORT-WAVE EXTRAPOLATION

275

h0
= ,
v =

r1
hA ,
2
2

r23
r1 r23
h, = 2 2 h2 A, .
2
3
2 3

(7.50)

Thus, in the low-frequency limit the internal degrees of freedom are determined by the external ones. Substituting (7.50) into (7.49) and keeping only
the principal terms, we arrive at the classical 2-D action functional (3.9).
Note also the similarity between this reduction and the reduction from Timoshenkos beams to Bernoulli-Eulers beams.

276

CHAPTER 7. ELASTIC SHELLS


Problems

1. Prove that the short-wave extrapolation of the average Lagrangian does


not depend on the choice of the normalization constant c.
2. Show that as 1/3, s2 4

100
3 2

and s3 1 +

52
.
3 2

3. Plot the coefficient s3 as the function of in the interval (0.44, 0.5).


4. Provide the similar short-wave extrapolation for elastic plates.

7.4

Dispersion of waves in plates

Long-wave asymptotics of the thickness branches. Let us apply the


theory of high-frequency vibrations to study the dispersion of waves in an
infinite plate of a thickness h, for which b = 0. We first study the dispersion of waves described by equations (7.36) and (7.37), which should be
asymptotically exact in the long-wave range. Take for example the branch
F (n) (flexural waves). Introducing the dimensionless variables
=

tcs
,
h

x
,
h

(7.51)

we rewrite equation (7.36) in the form


v| = 2 v + k1 2 v.
Seeking the solution of this equation in the form of a harmonic plane wave
v = aei(

1 )

we end up with the following dispersion relation


2 = 2 + k1 2 ,

(7.52)

where = 2n/e and


k1 =

16 tan(/2)
1

.
2
e

The dispersion relation for L (n) has exactly the same form, with = (2n+
1)/e and
1
16 cot(/2)
k1 = 2 +
.
e

Figure 7.4 shows the dispersion curves of two branches L (0) and F (1) for
= 0.31. Note that the coefficient k1 of L (0) is negative for this Poissons

7.4. DISPERSION OF WAVES IN PLATES

277

ratio so that the real dispersion curve for L (0) does not exist for > c .
It is also easy to check that the phase and group velocities are of opposite
signs in the long-wave range for L (0). Such wave motions carry energy in
one direction but appear to propagate in the other direction and are called
backward waves. This phenomenon plays an important role in posing the
conditions at infinity (principle of radiation) to select the unique solutions of
the boundary-value problems for semi-infinite plates.

Figure 7.4: Dispersion curves of L (0) and F (1) for = 0.31: 1L (0),
2F (1).
We now analyze the branch Fk (n). In terms of the dimensionless variables
(7.51) equation (7.37) reads

| = 2 + (k1 + 1)|
+ 2 .

(7.53)

We seek the solutions of (7.53) in the form


= a ei(

1 )

(7.54)

Substituting (7.54) into (7.53), one can see that there are two possible types
of waves corresponding to
1 = a1 ei(

1 )

i( 1 )

2 = a2 e

2 = 0 F-wave,
1 = 0 AS-wave.

For the F-waves the dispersion relation reads


2 = 2 + (k1 + 2)2 ,

(7.55)

278

CHAPTER 7. ELASTIC SHELLS

with = (2n + 1) and


k1 = 1 +

16e2 cot(/2)
,

while for the AS-waves (antisymmetric shear waves) we have


2 = 2 + 2 .

(7.56)

We also obtain exactly the same dispersion relations (7.55) and (7.56) for the
L-waves and SS-waves of the branch Lk (n), respectively, with = 2n and
k1 = 1

16e2 tan(/2)
.

Figure 7.5 shows the asymptotics of dispersion curves of two branches Lk (1)
and Fk (0) for = 0.31.

Figure 7.5: Dispersion curves of Lk (1) and Fk (0) for = 0.31.


Long-wave asymptotics according to 3-D elasticity. The dispersion
relation for waves in the plate according to 3-D elasticity (Rayleigh-Lamb
equation) has been derived in Section 3.3. For L-waves the equation (3.76)
can be rewritten in the form
p2
p1
p1
p2
(2 p22 )2 sin cos + 42 p1 p2 sin cos
= 0,
(7.57)
2
2
2
2
where
p1 =

e2 2 2 ,

p2 =

2 2 .

Note that if (, ) is a solution of (7.57), then (, ) are also its solutions.


Therefore it is enough to consider solutions of (7.57) in the first quadrant

7.4. DISPERSION OF WAVES IN PLATES

279

of the (, )-plane. Since we are interested in the high-frequency long-wave


range, we shall analyze (7.57) in the region III of this plane (see Figure 3.8).
Setting = 0 in (7.57), we see that the cut-off frequencies c are the roots
of the equation
ec
c
= 0.
sin cos
2
2
It implies that
(2n + 1)
c = 2n, or c =
.
e
The first series of roots correspond to the cut-off frequencies of the series Lk
from (7.19), the second one to the cut-off frequencies of the series L from
(7.18).
Consider the branch Lk (n). To study the asymptotics of the dispersion
curve near the cut-off frequency = 2n we introduce the notation
2 = 2 + y,

2 = x,

(7.58)

with x and y being small quantities. Expanding the left-hand side of the
equation (7.57) in the Taylor series of x and y and keeping only the principal
terms in accordance with Newtons rule, we obtain
4 cos

e
e

1
(y x) cos
+ 4xe 2 sin
cos = 0.
2 4
2
2
2

(7.59)

Solve this with respect to y to get


y = (1

16e tan(e/2)
)x,

which is the same as equation (7.55).


For the branch L (n) we introduce (7.58) with = (2n + 1)/e. After
performing the same operations as in the previous case we have
1
1
1

4 sin ( sin ) ( 2 y 2 x) + 4x sin cos = 0.


2 2
2 2

2
2
Solution of this equation with respect to y yields
y=(

1
16 cot(/2)
+
)x,
2
e

which is the same as equation (7.52).


Analogously, the asymptotic analysis of the Rayleigh-Lamb equation for
F-waves
p2
p1
p1
p2
(2 p22 )2 cos sin + 42 p1 p2 cos sin
= 0,
2
2
2
2

280

CHAPTER 7. ELASTIC SHELLS

in the long-wave range leads to the equations (7.52) and (7.55) for the corresponding branches.
In the above consideration we implicitly assume the value of e such that
cos(en) 6= 0. In the opposite case the coefficient at y in the approximate
dispersion equation (7.59) vanishes, and the equation (7.55) fails to provide
the true asymptotics for long waves. Consider, for definiteness, the branch
Lk (n) and introduce the new variables
= + y,

2 = x.

Expanding (7.57) in x and y and keeping their principal terms, we arrive at


4 cos

e e 2
e

sin
y + 4xe 2 sin
cos = 0,
2
2 4
2
2

yielding
=

4
.

(7.60)

Take, for instance, e = 0.5 ( = 1/3). One can see from (7.60) that the
group velocity cg = d/d of Lk (1) does not vanish at = 0, but is equal
to cg = 2/, and consequently, the wave train moves without deformation
in the long-wave range. It is also interesting to observe that, for = 1/3,
the cut-off frequency of the branch Lk (1) coincides with that of the branch
L (0). The orthogonality between these branches in the long-wave range is
no longer valid in this case, and the account of their interaction becomes
essential. We shall see later how the dispersion curves in the whole range of
wavelength look like.
Dispersion of waves in the whole range of wavelength
Flexural waves. We shall derive and study the dispersion equation for waves
in an infinite plate based on the 2-D theory of high-frequency vibrations
of plates, for which the 2-D equations of motion (7.8) break up into the
equations of longitudinal and flexural waves. Introducing the dimensionless
variables (7.51) and the new unknown function = h , we can write down
the equations for flexural waves in the form

+ 2 u),
u| = 12 (|
1

+ 2 ].
= 12 ( + u| ) + [(2 + 1)|
12

(7.61)

We seek solutions of these equations in the form of the harmonic waves


1
(u, 1 , 2 ) = (a, b, c)ei( ) ,

(7.62)

7.4. DISPERSION OF WAVES IN PLATES

281

where is the dimensionless wave number and the dimensionless frequency.


Substitute this formula into (7.61). The condition of non-triviality of the
solutions (7.62) yields two dispersion relations according to the two independent waves
1
2 = cei( ) ,

(u, 1 ) = (a, b)e

i( 1 )

u = 1 = 0 AS(0),
, 2 = 0 F (0) + Fk (0).

The dispersion relation for the branch AS(0) reads


2 = 12 +

1 2
.
12

(7.63)

According to 3-D elasticity the dispersion relation for waves of this type is
given by
2 = 12 + 2 .

(7.64)

Thus, the formula (7.63) correctly describes the cut-off frequency c = 1 (at
= 0), but not the curvature of the exact dispersion curve. This is in agreement with the derivation of the approximate 2-D theory. Note, however, that
1
= 16 ( 242 )2 0.9855. The
the error in (7.63) turns out to be small, because 12
approximate and exact dispersion curves are shown in Figure 7.6. If < 1 ,
the wave number becomes purely imaginary. Solutions associated with imaginary wave number decay exponentially with 1 and become important in the
problem of wave propagation in semi-infinite plates. Since the dispersion
curves are symmetric about the and axis, the negative half of the real
(Re, ) plane can be replaced by the positive half of the imaginary (Im, )
plane without losing any detailed information.
For the purely flexural waves (branches F (0) + Fk (0)) the following dispersion relation can be obtained from (7.61)
(12 2 2 )(

+1 2
+ 12 2 ) 2 14 2 = 0.
6

(7.65)

In the long-wave range (  1) the asymptotic analysis of (7.65) yields the


following formula
2 =

+1 4
+ O(6 ),
6

for the branch F (0), and


2 = 12 + (12 +

+1 2
) + O(4 ),
6

(7.66)

282

CHAPTER 7. ELASTIC SHELLS

Figure 7.6: Dispersion curves of AS-waves: a) 2-D theory: dashed line and
b) 3-D theory: solid line.
for the branch Fk (0). Comparing this with the long-wave asymptotics (3.81)
and (7.55) derived from 3-D elasticity, we can see that the low-frequency
branch F (0) is described by the 2-D theory asymptotically exactly up to
terms of the order O(6 ), while the high-frequency branch Fk (0) according
to the 2-D theory admits the error O(2 ), in agreement with the chosen
approximation. This means that the corresponding dispersion curve has the
same cut-off frequency, but different curvature at = 0 as compared with
the exact dispersion curve. Note, however, that the coefficients of 2 in (7.55)
and (7.66) do not differ much from each other. Taking for example = 1/3
we have
k1 + 2 = 1 +

16 cot(e/2)
3.546,

12 +

+1
3.791.
6

Thus, the relative error in the coefficient of 2 is about 7%.


In the short-wave range ( ) both the dispersion curves according
to (7.65) approach the asymptotes
r

+1
= 1 , =
,
6
in contrast to the classical 2-D theory. This property is characteristic for
hyperbolic systems. For = 1/3 the slopes of these lines are equal to
r

+1
1 0.9135 (cr /cs ),
1.7195 (1),
6
where the values in parentheses are obtained from 3-D elasticity, with cr
the Rayleigh velocity. The dispersion curves according to the 2-D and 3-D

7.4. DISPERSION OF WAVES IN PLATES

283

Figure 7.7: Dispersion curves of F-waves: a) 2-D classical theory: dotted


line, b) 2-D theory of high frequency vibrations: dashed line, and c) 3-D
theory: solid line.
theories are plotted in Figure 7.7 for = 0.31, where a qualitative agreement
between the 2-D theory of high-frequency vibrations and 3-D elasticity is
observed even in the short-wave range.
Longitudinal waves. In terms of the dimensionless variables (7.51) the 2-D
equations for longitudinal waves read
u| = (s1 + 1)u| + 2 u + r1 | ,

| = s2 2 r1 u| r23 v|
22 ,

v| = (s3 +

1)v|

+ v + r23 |

(7.67)
32 v .

Harmonic plane waves propagating in the 1 direction also fall into two
classes. For SS-waves (symmetric shear waves) of the type
u2 = aei(

1 )

v2 = bei(

1 )

u1 = = v1 = 0,

the following dispersion relations hold true


SS(0) : 2 = 2 ,
SS(1) : 2 = 32 + 2 .

(7.68)

Equations (7.68) coincide with the exact dispersion relations obtained from
3-D elasticity. The corresponding dispersion curves are plotted in Figure 7.8.

284

CHAPTER 7. ELASTIC SHELLS

Figure 7.8: Dispersion curves of SS-waves.


We now derive the dispersion relation for the longitudinal waves of the
second class
(u1 , , v1 ) = (a, b, c)ei(

1 )

u2 = v2 = 0.

Substitution of (7.69) into (7.67) leads to


m11
m12
0
a
m12 m22 m23 b = 0,
0
m23 m33
c

(7.69)

(7.70)

where
m11 = S1 2 2 , m12 = ir1 , S1 = s1 + 2,
m22 = S2 2 2 + 22 , m23 = ir23 , S2 = s2 ,
m33 = S3 2 2 + 32 , S3 = s3 + 2.

(7.71)

Equation (7.70) has non-trivial solutions if and only if its determinant vanishes
m11 m22 m33 + m212 m33 + m11 m223 = 0.

(7.72)

Replacing mij in (7.72) by their expressions in (7.71), we obtain the dispersion equation
(S1 2 2 )(S2 2 2 + 22 )(S3 2 2 + 32 )
2 2
(S3 2 2 + 32 )r12 2 (S1 2 2 )r23
= 0.

(7.73)

The cut-off frequencies are obtained by setting = 0 in (7.73). They are


equal to c = 0, c = 2 = /e and c = 3 = 2, which correspond to the

7.4. DISPERSION OF WAVES IN PLATES

285

cut-off frequencies of the three branches Lk (0), L (0) and Lk (1), respectively.
The asymptotics of the dispersion curves near the cut-off frequencies can be
derived by introducing the new variables 2 = 2c + y and 2 = x and
keeping in (7.73) the main terms of y and x in accordance with Newtons
rule. Standard calculations show that the asymptotic formulae (3.83), (7.52)
and (7.55) are also derivable from (7.73).

Figure 7.9: Dispersion curves of L-waves: a) 2-D classical theory: dotted line,
b) 2-D theory of high frequency vibrations: dashed line, and c) 3-D theory:
solid line.
Consider the special case = 1/3, for which the cut-off frequencies of the
branches L (0) and Lk (1) coincide. To derive the long-wave asymptotics of
these branches we introduce the variables = 2 + y and 2 = x. Expanding
(7.73) in y and x and keeping the main terms in accordance with Newtons
rule, we arrive at
2
424 y 2 + 22 r23
x = 0.

Taking into account that r23 = 8 for = 1/3, one can see that this equation
is nothing but the exact asymptotics (7.60). Thus, the asymptotic exactness
of (7.73) is proved for [0, 0.44).
In the short-wave range ( ) all the dispersion curves according to
(7.65) approach the asymptotes
=

S2 ,

p
S3 ,

p
S1 ,

286

CHAPTER 7. ELASTIC SHELLS

confirming the hyperbolicity of the 2-D equations (7.67). For = 1/3 the
slopes of the asymptotes are
p
p
p
S2 = 0.789 (cr /cs ),
S3 = 1.6602 (1),
S1 = 1.9521 (1),
where the values in parentheses are exact. The dispersion curves according to
the 2-D and 3-D theories are plotted in Figure 7.9 for = 0.31. When > 3
all the three branches according to the 2-D theory are real. For between
the two cut-off frequencies 2 and 3 the third branch forms a loop in the
imaginary plane. For (? , 2 ) the third branch becomes real again, but
its phase and group velocities are opposite. Finally, when < ? the second
and third branches are complex conjugate. Their projections onto the real
(Re, ) and imaginary (Im, ) planes are marked with r and i, respectively.
One can see that, in addition to the long-wave range, the 2-D theory of highfrequency vibrations describes satisfactorily the first complex branches of
the dispersion curves, which gives reason to expect a good prediction of the
exponentially decayed boundary layer by this theory.
Problems
1. Plot the dispersion curves according to (7.73), including the imaginary
and complex branches, for > 1/3.
2. Do the same as in problem 1 for = 1/3.

3. Plot the dispersion curves according to the Rayleigh-Lamb equation


(7.57), including the first complex branch, for = 1/3 and compare
with the result of the previous problem.
3. Prove that all the thickness branches according to 3-D elasticity approach asymptotically the line = from above as .

7.5

Frequency spectra of plates

Flexural vibrations. We first study in some detail the high-frequency


spectra of a circular plate of radius r. For its flexural vibrations the governing
equations are given by (7.61). Looking for solutions of the form
u( , ) = u( )ei ,

( , ) = ( )ei ,

with a frequency, we reduce equations (7.61) to

12 (|
+ 2 u) + 2 u = 0,
1

+ 2 ] + 2 = 0.
12 ( + u| ) + [(2 + 1)|
12

(7.74)

7.5. FREQUENCY SPECTRA OF PLATES

287

Now, using Helmholtzs decomposition theorem, we express u and in


terms of the following three scalar potentials
u = 1 + 2 ,

= a1 1| + a2 2| + .
. | ,

(7.75)

where a1 , a2 are constants. Substituting (7.75) into (7.74)2 and separating


the equations for 1 , 2 and , one can show that
2 + 12(2 12 ) = 0.

(7.76)

Further, equations (7.74) are fulfilled, provided 1 , 2 satisfy Helmholtzs


equations
2 1 + 21 1 = 0,

2 2 + 22 2 = 0,

(7.77)

where the constants a1 , a2 are chosen as follows


a1 =

2 12 21
,
12 21

a2 =

2 12 22
,
12 22

and, finally, 21 , 22 are the roots of the dispersion equation


(12 2 2 )(

+1 2
+ 12 2 ) 2 14 2 = 0.
6

The equations (7.76) and (7.77) can be solved by the separation of variables as in Section 3.4. Referring the plate to the polar co-ordinates %, ,
one can represent the general solutions by
(
(
cos n
cos n
1 = b1 Jn (1 %)
, 2 = b2 Jn (2 %)
,
(7.78)
sin n
sin n
and
(
sin n
= b3 Jn (3 %)
cos n

(7.79)

In the last equation 23 is defined as


23 = 12(2 12 ).
Now consider the traction-free edge. For flexural vibrations the boundary
conditions (7.9) reduce to
( + u| ) = 0,

( |
a + (|) ) = 0.

(7.80)

288

CHAPTER 7. ELASTIC SHELLS

Figure 7.10: Frequencies of axisymmetric flexural vibrations of plates with


free edges as functions of r/h for = 0.31.
Substitution of (7.75) in (7.80) leads to
1
2 1
+ (1 + a2 )
+
]|%=r = 0,
(7.81)
%
%
%
2 (a1 1 + a2 2 ) 1 2
1
[2 (a1 1 + a2 2 ) +
+
2
]|%=r = 0,
2
%
% % %
2 2 (a1 1 + a2 2 )
2 (a1 1 + a2 2 )
2
[
2
+ 2 2 2 ]|%=r = 0,
%
%
%

%
[(1 + a1 )

where r = r/h. Substituting for 1 , 2 and their expressions (7.78) and


(7.79), respectively, one obtains a system of three homogeneous linear equations in three unknowns b1 , b2 and b3 . The condition of vanishing determinant
yields the frequency equation for .
The frequency equation in the general case is rather complicated; therefore we shall for simplicity analyze the axisymmetric vibrations, for which
n = 0 and
1 = b1 J0 (1 %), 2 = b2 J0 (2 %), = 0.
In this case the third boundary condition of (7.81) is satisfied identically.
From the first two we obtain the following frequency equation
J1 (2 r)
J1 (1 r)
a2 22 [
( + 1)J0 (2 r)]
1
2 r
J1 (2 r)
J1 (1 r)

a1 21 [
( + 1)J0 (1 r)] = 0.
2
1 r

7.5. FREQUENCY SPECTRA OF PLATES

289

The first seven frequencies of this equation as functions of the ratio r = r/h
are computed for = 0.31. The spectra are shown in Figure 7.10. As r
increases, all the frequencies diminish and approach Poissons frequencies as
computed by equation (3.92) of the classical plate theory and shown by the
dashed lines in this figure.
Longitudinal vibrations. The longitudinal vibrations of the plate are
described by the equations (7.67). Looking for solutions of the form
u = u ( )ei ,

)ei ,
= (

v = v ( )ei ,

we reduce (7.67) to
(s1 + 1)
u| + 2 u + r1 | + 2 u = 0,
s2 2 r1 u r23 v + (2 2 ) = 0,
|

(s3 +

1)
v|

+ v + r23 | + (
2

2
2
3 )
v

(7.82)

= 0.

Again, making use of Helmholtzs decomposition theorem, we can express


u , and v in terms of five scalar potentials
u = 1| + 2| + 3| + .
. | ,
= a1 1 + a2 2 + a3 3 ,
v = b1 1| + b2 2| + b3 3| + .
. $| ,
where ai , bi are constants. In a similar manner as in the previous subsection
we can show that the equations (7.82) will be satisfied if i , and $ satisfy
Helmholtzs equations
2 i + 2i i = 0, i = 1, 2, 3,
2 + 2 = 0,
2 $ + (2 32 )$ = 0,

(7.83)

where 2i , i = 1, 2, 3, are the three roots of the equation (7.73) and the constants ai , bi are chosen as follows
S1 2i 2
,
r1
r23 ai
bi =
,
2
S3 i + 32 2
ai =

(7.84)

with S1 , S3 taken from (7.71). The solutions of (7.83) are given in the form
(
cos n
i = ci Jn (i %)
, i = 1, 2, 3,
(7.85)
sin n

290

CHAPTER 7. ELASTIC SHELLS

and
(
sin n
= c4 Jn (%)
cos n

(
sin n
$ = c5 Jn (5 %)
cos n

(7.86)

In the last equation 25 is defined as


25 = 2 32 .
For the free edge the following boundary conditions should be fulfilled
= 0,
s1 u| + 2
u(|) + r1
s2 | r3 v = 0,
= 0.
s3 v + 2
v (|) + r2
|

In terms of i , and $ these conditions are transformed to


X
1
1 2
2 X

)
+
r

+
2(
ai i ]|%=r = 0,
1
i
%2
% % %2
2 2 X
2 X
2
[
i 2
i + 2 2 2 ]|%=r = 0,
% %
%
%
X
X

r3 $
[s2
]|%=r = 0,
(7.87)
i + r3
i +
%
%
%
X
X
1 $
2 X
1 2$
[s3 2
2
) + r2
bi i + 2 2
bi i + 2(
ai i ]|%=r = 0,
%
% % %
2 2 X
2 X
2$
[
bi i 2
bi i + 2 $ 2 2 ]|%=r = 0,
% %
%
%
[s1 2

i + 2

where all the sums are over i running from 1 to 3. Substituting (7.85) and
(7.86) into these conditions, we obtain the system of five homogeneous linear
equations with respect to the five unknown coefficients c1 , c2 , c3 , c4 , c5 , the
determinantal equation of which determines the frequencies of vibrations.
We further restrict ourselves to the axisymmetric vibrations, for which
n = 0 and
i = ci J0 (i %), i = 1, 2, 3, = $ = 0.
In this case the second and the fifth boundary conditions of (7.87) are satisfied
identically. From the remaining ones we can obtain the following frequency
equation
det Cij = 0,

(7.88)

7.5. FREQUENCY SPECTRA OF PLATES

291

where
i
J1 (i r) 2 J0 (i r),
r
C2i = (S2 ai + r3 bi )i J1 (i r),
i
C3i = (r2 ai S3 bi 2i )J0 (i r) + 2bi J1 (i r),
r
C1i = 2

i = 1, 2, 3,
i = 1, 2, 3,
i = 1, 2, 3.

The evaluation of the determinant in (7.88) is simplified by the fact that


it is an even analytic function of i , the roots of the equation (7.73). The
latter ones, for different ranges of , are seen in Figure 7.9. For < ? one
of the 2i is real and the other two are complex conjugate. If we denote by
23 the complex conjugate of 22 , then the terms of the third column of the
determinant in (7.88) are the complex conjugates of the terms of the second
column; i.e.,
Cj3 = Cj2 = Rj2 iIj2 ,
where Rj2 and Ij2 are the real and imaginary parts of Cj2 , respectively.
Hence, adding and subtracting the two columns one obtains


C11 R12 I12


det Cij = 4i C21 R22 I22 ,
C31 R32 I32
i.e., the characteristic determinant is purely imaginary. The frequencies are
plotted against the radius-to-thickness ratio r in Figure 7.11 for = 0.31.
As r increases all the frequencies diminish and approach the frequencies
computed by the equation (3.104) of the classical plate theory (the dashed
lines in Figure 7.11). As the frequencies increase, the typical terrace-like
structure of the high-frequency spectrum begins to develop at a frequency,
whose asymptotic value, as r , is e 4.56. This frequency is lower
than the first cut-off frequency of the thickness-stretch mode given by =
2 = /e. Accordingly, there is an important difference between longitudinal
and flexural vibrations of plates. In the latter, the asymptotic frequency of
the lowest terrace-like structure is always the cut-off frequency = of the
lowest thickness-shear mode.
We now examine the distribution of the displacements in the plate for the
frequencies near the terrace-like structure. Computations of u = u% , and
v = v% for r = 2.65, e = 4.541 (the point lying on the third terrace) show
that the amplitudes of displacements are large near the boundary of the plate
and decrease rapidly toward the centre (see Figure 7.12). The amplitude of
at the boundary is the largest one among all the three functions, and it
is nearly three times larger than that in the centre of the plate. This phenomenon is called edge resonance [41]. The frequency of the edge resonance

292

CHAPTER 7. ELASTIC SHELLS

Figure 7.11: Frequencies of axisymmetric longitudinal vibrations of plates


with free edges as functions of r/h for = 0.31.
computed by 3-D elasticity is equal to 4.66 [18]. Mindlin-Medicks theory
gives e = 4.14 [41]. The simple explanation of this phenomenon can be
given by studying the reflection of waves in a semi-infinite strip.
Reflection of waves in a semi-infinite strip. Consider the wave propagation in a semi-infinite strip bounded by the free edge at x1 = 0. Assuming
that u2 = v2 = 0 and u1 , and v1 depend only on 1 , we obtain from (7.82)
the following equations
S1 u1|11 + r1 |1 + 2 u1 = 0,
S2 |11 r1 u1|1 r23 v1|1 + (2 22 ) = 0,
S3 v1|11 + r23 |1 + (2 32 )
v1 = 0,

(7.89)

and the free edge boundary conditions at 1 = 0


S1 u1|1 + r1 = 0,
S2 |1 + r3 v1 = 0,
S3 v1|1 + r2 = 0.

(7.90)

We need to find a non-trivial solution of (7.89) and (7.90) which is bounded


as 1 . As in the previous case, u1 , and v1 can be expressed through

7.5. FREQUENCY SPECTRA OF PLATES

293

Figure 7.12: Distributions of u, and v at the frequency of edge resonance


(
r = 2.65, e = 4.541).
the three potentials i as follows
u1 = 1|1 + 2|1 + 3|1 ,
= a1 1 + a2 2 + a3 3 ,
v1 = b1 1|1 + b2 2|1 + b3 3|1 ,

(7.91)

where i satisfy the equations


i|11 + 2i i = 0,

i = 1, 2, 3,

(7.92)

with 2i , i = 1, 2, 3, being the roots of the equation (7.73), and the constants
ai , bi chosen according to (7.84). Formula (7.91) is convenient in the sense
that it enables one to separate waves with different i . Let the frequency
< ? , i.e. the cubic equation (7.73) has one positive real root 21 and two
complex conjugate roots 22 and 23 . The solutions of (7.92) can be written
as
1

1 = ei1 + c1 ei1 ,
1

2 = c2 ei2 ,

3 = c3 ei3 ,

where c1 , c2 , c3 are unknown constants. Without limiting generality we can


1
set the coefficient of the incident wave ei1 propagating toward 1 = 0 equal
to 1. In order for the solution to be bounded at infinity, 2 and 3 should be
taken from the lower half of the complex plane: Im2 , Im3 < 0.

294

CHAPTER 7. ELASTIC SHELLS

Having the potentials, we can now express u1 , and v1 through them


1

u1 = i1 (ei1 c1 ei1 ) i2 c2 ei2 i3 c3 ei3 ,


1
1
1
1
= a1 (ei1 + c1 ei1 ) + a2 c2 ei2 + a3 c3 ei3 ,
1

v1 = b1 i1 (ei1 c1 ei1 ) b2 i2 c2 ei2 b3 i3 c3 ei3 .


Substituting these formulae into the boundary conditions (7.90) we obtain
the system of three linear equations with respect to ci
c1 + c2 + c3 = 1,
C21 c1 + C22 c2 + C23 c3 = C21 ,
C31 c1 + C32 c2 + C33 c3 = C31 ,

(7.93)

where the components C2j , C3j are given by


C2j = j (S2 aj + r3 bj ),
C3j = S3 bj 2j r2 aj ,

j = 1, 2, 3,
j = 1, 2, 3.

Figure 7.13: Graphs of Re(c1 ) and Im(c1 ) versus .


From the linear system (7.93) one obtains three amplitudes of the reflected
waves, with two complex waves corresponding to the motion localized near
the free edge of the strip. Since 1 is real and 3 =
2 , one can easily check
that
|c1 | = 1, |c2 | = |c3 |.
The complex value of c1 implies that the reflected wave corresponding to 1
has a phase difference with respect to the incident wave given by
tan = Im(c1 )/Re(c1 ).

7.5. FREQUENCY SPECTRA OF PLATES

295

Figure 7.14: Graph of |c2 | versus .


In Figure 7.13 the graphs of Re(c1 ) and Im(c1 ) as functions of are shown
(Poissons ratio is equal to 0.31). It is seen that for  e the phase
difference between the incident and reflected waves is nearly . At = e
4.56 we have Re(c1 ) = 1 and Im(c1 ) = 0, i.e., the phase difference vanishes
at this frequency, regarded as the frequency of the edge resonance. In Figure
7.14 one can see the graph of |c2 | as a function of . At this frequency
|c2 | reaches its maximum. If is increased beyond e the amplitude |c2 |
decreases rapidly to very small values and the phase difference changes sign
and increases rapidly from 0 to nearly . In Figure 7.15 the dependence of
the edge resonance frequency on Poissons ratio (0 0.35) is shown,
where small circles correspond to the edge resonance frequency computed by
3-D elasticity [18].

Figure 7.15: Edge resonance frequency e versus Poissons ratio .


An important feature, observable in Figure 7.14, is that the amplitude

296

CHAPTER 7. ELASTIC SHELLS

|c2 | passes through a finite maximum and does not reach infinity. This means
that the large edge-deformation is coupled with a non-decaying branch for
that particular value of . The special case = 0, where a pure edge mode
exists, is worthy of mention. Since = r1 = r2 = 0, the boundary-value
problem (7.89) and (7.90) breaks into two problems. The first problem
S1 u1|11 + 2 u1 = 0,

u1|1 = 0 at 1 = 0

is classical and has no exponentially decayed solution. The second problem,


consisting of the equations
S2 |11 + r3 v1|1 + (2 22 ) = 0,
S3 v1|11 r3 |1 + (2 32 )
v1 = 0,

(7.94)

and the free edge boundary conditions at 1 = 0


S2 |1 + r3 v1 = 0,
v1|1 = 0,

(7.95)

has such solutions, which can be determined analytically. Indeed, it is easy


to see that solutions of (7.94) can be presented in the form
= 1 + 2 ,
v1 = b1 1|1 + b2 2|1 ,
where i , i = 1, 2, satisfy (7.92), with 2i being the roots of the equation
(S2 2i 2 + 22 )(S3 2i 2 + 32 ) r3 2i = 0,

(7.96)

and with bi given by


bi = r3 /(S3 2i 2 + 32 ).
For < ? the quadratic equation (7.96) has two complex conjugate
roots. Therefore the solutions bounded at infinity are given by
1

1 = c1 ei1 ,

2 = c2 ei2 ,

where 1 , 2 should lie in the upper half of the complex plane (Re,Im).
Substituting the solution into the free edge boundary conditions (7.95) and
equating the determinant to zero, we obtain the equation determining the
frequency of the edge mode
(S2 + r3 b1 )2 b2 = (S2 + r3 b2 )1 b1 .

(7.97)

7.6. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

297

Figure 7.16: Deformed shape of the trip at the edge resonance frequency e .
After some algebra we can transform it to
S2 (S2 S3 )4e + S2 (S3 22 S2 32 + 2r32 S2 32 + S3 32 )2e
+ [r34 S2 32 (S3 22 S2 32 + 2r32 )] = 0.

(7.98)

For = 0 equation (7.98) gives two roots 3.9387 and 8.8812. The first root
e = 3.9387 corresponds to the frequency of the edge mode.7 In Figure 7.16
the deformed shape of the strip is shown. One can see that the solution decays exponentially, and the width of the boundary layer, where the intensive
motion of the edge can be observed, is of the order h.
Problems
1. Study the frequency spectra of the flexural vibrations of circular plates
for n = 1.
2. Study the frequency spectra of the longitudinal vibrations of circular
plates for n = 1. Does the edge resonance exist? If yes, find the
frequency corresponding to this mode of vibrations.
3. Determine the edge resonance frequency of the circular plate for = 0
and compare it with that of the strip. Does an exponentially decayed
solution exist? If so, find it.

7.6

Dispersion of waves in cylindrical shells

2-D equations of motion. In this section we apply the theory of highfrequency vibrations to study the dispersion of waves in an infinite circular
7

According to 3-D elasticity, e = 3.9584 [18].

298

CHAPTER 7. ELASTIC SHELLS

cylindrical shell of the thickness h and the radius R. We denote by x1 the


axial and by x2 = R the circumferential co-ordinate of the shell middle
surface, respectively (Figure 3.13). The components of the first and second
quadratic forms of the shell middle surface are given by equations (3.105).
Due to the fact that the metrics of the surface is the Kronecker delta, the
operations of raising or lowering of indices do not affect the components
of tensors at all, so we are allowed to lower all indices. Besides, since the

Christoffel symbols
vanish, the covariant derivatives coincide with the
corresponding partial derivatives. It is easy to compute the measures of
extension as given by (7.4)1
1
A12 = (u1,2 + u2,1 ),
2

A11 = u1,1 ,

A22 = u2,2 +

u
.
R

The measures of bending given by (7.4)2 in the chosen co-ordinate system


are found as follows
11 = 1,1 , 22 = 2,2 ,
1
1
12 = (1,2 + 2,1 )
(u2,1 u1,2 ).
2
4R
We now write down the differential equations of motion of the cylindrical
shell in components. According to (7.8)
h
u1 = t11,1 + t12,2 ,
1
h
u2 = t21,1 + t22,2 + q2 ,
R
1
h
u = q1,1 + q2,2 t22 ,
R
h3 1 = q1 m11,1 m12,2 ,
h3 2 = q2 m21,1 m22,2 ,
h = h[s2 2 r1 h1 (u1,1 + u2,2 +

(7.99)
1
u)
R

r23 h1 (v1,1 + v2,2 ) f2 h2 ],


h
v1 = h[(s3 + 1)(v1,11 + v2,21 ) + 2 v1 + r23 h1 ,1 f31 h2 v1 ],
h
v2 = h[(s3 + 1)(v1,12 + v2,22 ) + 2 v2 + r23 h1 ,2 f32 h2 v2 ],
where
t11 = n11 ,
t12

t22 = n22 ,
m12
m12
, t21 = n12
,
= n12 +
2R
2R

7.6. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

299

and the concise notation has been introduced


f31 = f311 ,

f32 = f322 .

The equations (7.99) should be complemented by the constitutive equations expressing the membrane stresses n , the bending moments m and
q through the measures of extension A , bending and rotations + .
These constitutive equations, in component form, are given by
r1
u
n11 = h[s1 (u1,1 + u2,2 + ) + 2u1,1 + ],
R
h
u
r1
u
n22 = h[s1 (u1,1 + u2,2 + ) + 2u2,2 + 2 + ],
R
R
h
1
n12 = 2h (u1,2 + u2,1 ),
2
3
h
m11 =
[(1,1 + 2,2 ) 1,1 ],
6
h3
m22 =
[(1,1 + 2,2 ) 2,2 ],
6
h3 1
1
m12 =
[ (1,2 + 2,1 )
(u2,1 u1,2 )].
6
2
4R
q1 = hf11 (1 + u,1 ),
u2
q2 = hf12 (2 + u,2 ),
R

(7.100)

where
15 2 2
h /R ),
4
3
= ( 2 + h2 /R2 ).
4

f11 = f111 = ( 2 +
f12 = f122

Substituting these equations into (7.99), we obtain the 2-D equations of


motions in terms of the unknown functions u , , v , u, . Introducing the
quantities
=

tcs
,
h

x
,
h

= h ,

h? =

h
,
R

(7.101)

we transform the equations of motion to the more convenient dimensionless

300

CHAPTER 7. ELASTIC SHELLS

form
u1| = [s1 (u1|1 + u2|2 + h? u) + 2u1|1 + r1 ]|1 + (u1|2 + u2|1 )|2
h? 1
h?
+ [ (1|2 + 2|1 ) (u2|1 u1|2 )]|2 ,
12 2
4
h?
h? 1
u2| = (u1|2 + u2|1 )|1 [ (1|2 + 2|1 ) (u2|1 u1|2 )]|1
12 2
4
+[s1 (u1|1 + u2|2 + h? u) + 2u2|2 + 2h? u + r1 ]|2 + h? f12 (2 + u|2 h? u2 ),
u| = f11 (1 + u|1 )|1 + f12 (2 + u|2 h? u2 )|2
h? [s1 (u1|1 + u2|2 + h? u) + 2u2|2 + 2h? u + r1 ],
1
1| = f11 (1 + u|1 ) [(1|1 + 2|2 ) 1|1 ]|1
6
h?
1 1
(7.102)
[ (1|2 + 2|1 ) (u2|1 u1|2 )]|2 ,
6 2
4
h?
1 1
2| = f12 (2 + u|2 h? u2 ) [ (1|2 + 2|1 ) (u2|1 u1|2 )]|1
6 2
4
1
[(1|1 + 2|2 ) 2|2 ]|2 ,
6
| = s2 (|11 + |22 ) r1 (u1|1 + u2|2 + h? u)
r23 (v1|1 + v2|2 ) f2 ,
v1| = (s3 + 1)(v1|11 + v2|21 ) + v1|11 + v1|22 + r23 |1 f31 v1 ,
v2| = (s3 + 1)(v1|12 + v2|22 ) + v2|11 + v2|22 + r23 |2 f32 v2 ,
where
15 2
h,
4 ?
3
= 4 2 + h2? .
4

f31 = f311 = 4 2 +
f32 = f322

Dispersion relation. We seek solutions of the wave equations (7.102) in


the form
(u1 , 1 , v1 ) = (w1 , w4 , w7 ) cos 2 sin( 1 ),
(u2 , 2 , v2 ) = (w2 , w5 , w8 ) sin 2 cos( 1 ),
(u, ) = (w3 , w6 ) cos 2 cos( 1 ),

(7.103)

where w1 , . . . , w8 are the unknown constants, and the dimensionless


longitudinal wave number and frequency, respectively, and is determined
through n, the number of waves around the circumferential, by
= nh? ,

n = 0, 1, 2, . . .

7.6. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

301

The periodic functions of 2 used in (7.103) guarantee that the displacements


are continuous with respect to 2 . Besides, the alternation of the sine and
cosine functions is required in order to make the trigonometric terms factor
out of the differential equations determining u , , v , u, . Substituting
(7.103) into (7.102) and eliminating the common factors, we arrive at the
following system of linear homogeneous equations
8
X

Hij (, )wj = 0,

i = 1, . . . , 8,

(7.104)

j=1

where Hij is an 8 8 matrix whose elements are real functions of and


. One can check directly that the matrix Hij is symmetric. Its diagonal
elements are given by
H11 = (s1 + 2)2 2

h2? 2
+ 2 ,
48

h2? 2
(s1 + 2) 2 h2? f12 + 2 ,
48
H33 = f11 2 f12 2 h2? (s1 + 2) + 2 ,
+1 2
1
H44 = f11
2 + 2 ,
6
12
1 2 +1 2
+ 2 ,
H55 = f12
12
6
H66 = s2 (2 + 2 ) f2 + 2 ,
H77 = (s3 + 2)2 2 f31 + 2 ,
H88 = (s3 + 2) 2 2 f32 + 2 ,

H22 = 2

while the out-of-diagonal elements are given by


h2?
), H13 = s1 h? ,
48
h?
h?
H14 = 2 , H15 = , H16 = r1 ,
24
24
h?
H23 = (s1 + 2 + f12 )h? , H24 = ,
24
h? 2
H25 = + h? f12 , H26 = r1 ,
24
H34 = f11 , H35 = f12 , H36 = h? r1 ,
H12 = (s1 + 1

(7.105)

302

CHAPTER 7. ELASTIC SHELLS


2 + 1
, H67 = r23 ,
12
H68 = r23 , H78 = (s3 + 1),
H17 = H18 = H27 = H28 = H37 = H38
= H47 = H48 = H56 = H57 = H58 = 0.
H45 =

= H46

(7.106)

The system (7.104) has non-trivial solutions if and only if the determinant
of Hij vanishes
G = det Hij = 0.

(7.107)

This is the dispersion relation for the waves (7.103) propagating in the cylindrical shell. One can see that for every fixed n and every real and fixed
the determinant in (7.107) is a polynom of eighth degree with respect to 2
giving eight values of in the upper half of the , -plane.
Asymptotics of the cut-off frequencies. Before studying the dispersion
of waves in the whole range of wavelength, we first consider the asymptotics
of the cut-off frequencies of high-frequency branches and then compare them
with those obtained from 3-D elasticity. Consider for simplicity the case
n = 0. Since the cut-off frequencies are found by the condition = 0, it
is easy to see that there are five cut-off frequencies of the high-frequency
branches. For the branch Fk (0) the cut-off frequencies are given by
15
f11
= 2 + h2? ,

4
3
f12
= 2 + h2? ,
2c =

2c =

1 6= 0,
(7.108)
2 6= 0.

For the branch L (0) we have

1
2c = f2 = ( )2 + h2? (4 2 ),
e
4e

6= 0.

(7.109)

Finally, for the branch Lk (1) the frequencies are given by


15 2
h,
4 ?
3
= 4 2 + h2? ,
4

2c = f31 = 4 2 +
2c = f32

v1 6= 0,
(7.110)
v2 6= 0.

We want to show that these asymptotic formulae confine to those obtained


by 3-D elasticity. Indeed, with n = 0 and = 0 the determinant in the

7.6. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

303

dispersion relation (3.125) according to 3-D elasticity breaks into three determinants
det Cij = D1 D3 D4 = 0,
where D1 is given by (3.127)1 and


C35 C36
,
D3 =
C65 C66



C11 C12
.
D4 =
C41 C42

The equation D1 = 0 determines the cut-off frequencies of the torsional


waves. It can also be written in the form (cf. (3.129))
J2 ()Y2 () Y2 ()J2 () = 0,

(7.111)

where = 1/h? +1/2, = 1/h? 1/2. For thin cylindrical shells h?  1, and
under the assumption of non-zero it is seen that  1 and  1. Using
the Hankel-Kirchhoff asymptotic approximations of the Bessel functions for
large argument |z|
r
2

15

[ cos(z ) +
sin(z )],
J2 (z)
z
4
8z
4
r

15

2
[ sin(z )
cos(z )],
Y2 (z)
z
4
8z
4
we transform the equation (7.111) to
sin

15
cos = 0.
8

This leads to the formulae (7.108)2 and (7.110)2 . In a similar manner, from
the equation D3 = 0, which is equivalent to
J1 ()Y1 () Y1 ()J1 () = 0,
we can derive the asymptotic formulae (7.108)1 and (7.110)1 . We now turn
to the equation D4 = 0 which describes vibrations with w 6= 0. It can be
written in the form
1
1
)J0 (e)][Y2 (e) + (1 2 )J0 (e)]
2
e
e
1
1
[J2 (e) + (1 2 )J0 (e)][Y2 (e) + (1 2 )J0 (e)] = 0. (7.112)
e
e

[J2 (e) + (1

304

CHAPTER 7. ELASTIC SHELLS

Again, using the asymptotic formulae of the Bessel functions J2 , J0 and Y2 , Y0


for large arguments, we reduce (7.112) to the following asymptotically equivalent equation
1
16 1/e2
cos e = 0.
sin
e

e2
8e
The asymptotics of one of the cut-off frequencies of this equation coincides
with (7.109).
Dispersion of waves in the whole range of wavelength
Axisymmetric waves. For waves independent of the angular co-ordinate 2
(n = 0), the determinant in (7.107) breaks into the product of three determinants, so that
G = G1 G2 G3 = 0,
where

G1 = H88 ,



H22 H25
,
G2 =
H52 H55


H11

H13

G3 = 0
H16

0

H13 0
H33 H34
H34 H44
H36 0
0
0


H16 0
H36 0
0
0 ,
H66 H67
H67 H77

and the terms Hij are given by (7.105) and (7.106) with n = 0. The equation
G1 G2 = 0 describes the dispersion of the torsional waves, for which
u1 = 1 = v1 = u = = 0.
In Figure 7.17 the dispersion curves of the torsional waves according to
the 2-D (dashed lines) and 3-D theories (solid lines, equation (3.129)) are
shown. The parameters chosen for the numerical calculations are equal to
= 0.31,

h/R = 1/10.

On can see that, up to the third branch, the dispersion curves according to
the 2-D and 3-D theories are practically identical, even for very short waves.
The dimensionless phase velocity for all branches decreases monotonically
from infinity to 1 as . At the same time the dimensionless group
velocity increases monotonically from zero to 1.
The equation G3 = 0 describes the dispersion of the AR-waves, with
u2 = 2 = v2 = 0.

7.6. DISPERSION OF WAVES IN CYLINDRICAL SHELLS

305

Figure 7.17: Dispersion curves of torsional waves in cylindrical shells: a) 2-D


theory: dashed line, and b) 3-D theory: solid line.
It leads to five dispersion curves, which, for = 0.31, h/R = 1/10, are shown
in Figure 7.18 (dashed lines) together with those computed by 3-D elasticity
according to the equation (3.130). In the long-wave range ( . 1.5) we see
that, up to the fifth branch, both curves are practically identical. Even in
the short-wave range a qualitatively good agreement between the 2-D and
3-D dispersion curves is observed.
Nonaxisymmetric waves. In this case all branches of waves are coupled, and
the dispersion curves according to the 2-D theory should be calculated from
(7.107). Taking for example
= 0.31,

h/R = 1/10,

n = 1,

we plot the dispersion curves in Figure 7.19 (dashed lines). The lowest branch
for n = 1 corresponds to the flexural waves whose frequency approaches
zero as the wave number tends to zero. The second and third branches are
identified as the lowest longitudinal shear branch and the ring-extensional
(breathing) branch, respectively. All branches higher than the third are
referred to as thickness branches. To be able to compare them with the analogous results from the three-dimensional theory, we show also the dispersion
curves obtained from Gazis dispersion relation (3.125) (the solid lines in
Figure 7.19). One can see that, up to the eighth branch, both curves are
almost identical in the long-wave range. Moreover, even in the short-wave
range there is a qualitatively good agreement between them.

306

CHAPTER 7. ELASTIC SHELLS

Figure 7.18: Dispersion curves of AR-waves in cylindrical shells: a) 2-D


theory: dashed line, and b) 3-D theory: solid line.
Problems
1. Plot the dispersion curves of waves for n = 2 in a closed circular cylindrical shell with = 0.31, h/R = 0.1.
2. Compare the dispersion curves for n = 1, 2.

3. Plot the first eight branches of the dispersion curves for n = 2 according
to Gazis equation (3.125) and compare them with the result of the first
problem.
4. Study the curvatures of the dispersion curves near the cut-off frequencies from the dispersion equation (7.107).

5. Compare the result of the previous problem with that obtained from
Gazis equation.

7.7

Frequency spectra of cylindrical shells

High-frequency approximation. Let a finite closed circular cylindrical


shell of the thickness h, the radius R, and the length 2L be referred to the
same co-ordinates as in Section 3.5. The 2-D equations of its motion are
given in (7.102). Although these equations appear rather complicated, the
high-frequency spectrum of the cylindrical shell turns out to be as simple

7.7. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

307

Figure 7.19: Dispersion curves of waves in cylindrical shells for n = 1: a)


2-D theory: dashed line, and b) 3-D theory: solid line.
as that of a strip. Before showing this, we note that by choosing the dimensionless variables (7.101), we try to examine the high-frequency spectra
under, say, the low-resolution microscope. The alternative dimensionless
variables chosen in (3.108) would then correspond to the high-resolution
microscope. In the matrix form equations (7.102) can be written as follows
Lu = 0,

(7.113)

where u correspond to the generalized displacements


uT = (u1 , u2 , u, 1 , 2 , , v1 , v2 ),
and L is a matrix differential operator. It is easy to see that
L = L0 + h? L1 ,

(7.114)

where L0 is the operator according to the theory of high-frequency vibrations


of elastic plates and L1 is a correction operator.
Analogously, the traction-free boundary conditions (7.9) are also decomposed into the sum of the principal terms corresponding to the plate theory
and the correction terms of order h? of smallness. Thus, if we admit the error
h? in determining the eigenfrequencies ,8 the terms of order h? in (7.114)
and in the boundary conditions can be neglected in the first step. Seeking
8

This means the error of order 1 for the dimensionless frequencies = R/cs .

308

CHAPTER 7. ELASTIC SHELLS

for simplicity the axisymmetric solution with u2 = 2 = v2 = 0 and with


the remaining functions depending only on 1 , two uncoupled systems are
obtained, namely, the equations
u1| = S1 u1|11 + r1 |1 ,
| = S2 |11 r1 u1|1 r23 v1|1 22 ,
v1| = S3 v1|11 + r23 |1 32 v1 ,
and boundary conditions
S1 u1|1 + r1 = 0,
S2 |1 + r3 v1 = 0,
S3 v1|1 + r2 = 0,
for the longitudinal vibrations, and the equations
u| = 12 (1|1 + u|11 ),
1
1| = 12 (1 + u|1 ) + ( + 1)1|11 ,
6
and boundary conditions
1 + u|1 = 0,
1|1 = 0,
for the flexural vibrations. The frequency equations are identical with those
for the elastic strip of length 2L and can be obtained in a standard manner.
In Figure 7.20 the eigenfrequencies of these systems are plotted against
the dimensionless length l = L/h of the shell. The intersection points of the
two sets of curves correspond to the multiple frequencies in the degenerate
case h? = 0. Now let h? be a small number differing from zero. Taking terms
of the oder h? in (7.113) and in the boundary conditions into account, we
obtain the coupled system with a small coupling between u1 , , v1 and u, 1 .
According to the theory of pertubations [45] the multiplicity of frequencies
should disappear. Obviously, the solid lines in Figure 7.20 change to the
dashed lines, which, for the typical intersection point, look like that shown
in Figure 7.21.
Thus, the high-frequency spectra of cylindrical shells (in the axisymmetric case) differ from those of strips only near intersection points by the mode
switching effect.9 Particularly, the edge resonance vibrations are also observed in semi-infinite cylindrical shells. For = 0.31 the edge resonance
frequency is equal to e = 4.56, which is the same as that of the strip.
9

One sees that this method does not work in the low-frequency range, since the error
is comparable with the eigenfrequencies.

7.7. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

309

Figure 7.20: Frequency spectra of elastic strips with free edges ( = 0.31).

Figure 7.21: Small perturbation of the frequency spectra.


Shells with shear diaphragms at both edges. It turns out that the
boundary conditions (7.10) posed at both edges of the cylindrical shell permit the simple solution of the equations of high-frequency vibrations. The
conditions of this type correspond to the shear diaphragm, the physical meaning of which has been explained in Section 3.6. By analogy we construct the
solution in the form
(u1 , 1 , v1 ) = (a1 , a4 , a7 ) cos 1 cos nei ,
(u2 , 2 , v2 ) = (a2 , a5 , a8 ) sin 1 sin nei ,
(u, ) = (a3 , a6 ) sin 1 cos nei ,
and choosing
m =

m
,
2l

(7.115)

the boundary conditions (7.10) are readily seen to be satisfied exactly. The
eigenfrequencies are determined by the dispersion equation (7.107), where

310

CHAPTER 7. ELASTIC SHELLS

Figure 7.22: Frequencies of the cylindrical shell with the shear diaphragms
at both edges (n = 1, = 0.31, h/R = 0.1): a) 2-D theory: dashed line, and
b) 3-D theory: solid line.
is replaced by m according to (7.115). Typical nodal patterns are shown in
Figure 3.22.
As was mentioned in Section 3.6, the same problem can be solved exactly
within the framework of 3-D elasticity. The three-dimensional boundary
conditions corresponding to this situation read
w2 = w3 = 0,

11 = 0,

at 1 = l.

The solution is given in the form of the standing waves (see formulae (3.145)).
The frequencies are determined from Gazis dispersion equation (3.125) with
replaced by m ; that means that they are determined by the points of
intersection of the dispersion curves with the equidistant vertical lines. The
frequencies according to the 2-D and 3-D theories are plotted in Figure 7.22
for n = 1, = 0.31, h/R = 0.1. The 2-D theory of high-frequency vibrations
works very well in both the low- and high-frequency long-wave ranges. One
observes a qualitatively good agreement between the theories even in the
short-wave range.
Problems
1. Determine the high-frequency spectra of the closed cylindrical shell
with the free edges for n = 1.

7.7. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

311

2. Determine the frequency spectrum of the closed cylindrical shell with


shear diaphragms at both edges for n = 2.

3. Compare the frequency spectrum in the previous problem with that


obtained from 3-D elasticity.

312

CHAPTER 7. ELASTIC SHELLS

Chapter 8
Elastic rods
8.1

One-dimensional equations

Variational principle. As was said in the Introduction, Poissons reduction


of the 3-D equations of elasticity to 1-D equations of low-frequency vibrations of rods has been the subject of many revisions and extensions to higher
frequencies. In this chapter we highlight the application of the variationalasymptotic method in deriving a one-dimensional theory of elastic rods that
describes their vibrations in a wide range of frequencies. Taking the variational approach as the basis, we postulate that the vibrations of the rod
in the absence of external body forces and surface tractions should occur in
accordance with the following variational principle
Z t1 Z L
J =
( ) dx dt = 0,
(8.1)
t0

where L is the length and and are the 1-D kinetic and strain energy
density of the rod, respectively. The functional (8.1) is called the 1-D action
functional. Assuming, for simplicity, that the rod is straight and untwisted
in the natural state, and its cross section is centrally symmetric, we shall
construct for it the average Lagrangians of high-frequency longitudinal and
flexural vibrations.
Longitudinal vibrations. In the classical theory of low-frequency longitudinal
vibrations of naturally straight and untwisted rods the following formulae
hold true
1
1
= h2 u 2 , = Eh2 (u0 )2 ,
2
2
where u(x, t) is the longitudinal displacement averaged over the cross section (up to a normalization constant). It is natural to expect that, as the
313

314

CHAPTER 8. ELASTIC RODS

frequencies increase, some internal degrees of freedom corresponding to the


branches of the rod thickness vibrations will become equally essential, so
that they should be included, apart from u, as additional unknown functions
in the 1-D kinetic and strain energy densities. Within this approach the
crucial point turns out to be the definition of the most essential degrees of
freedom and the specification of the 1-D kinetic and strain energy densities
depending on these degrees of freedom. The essence of the proposed theory
of high-frequency longitudinal vibrations of elastic rods can be expressed by
the following formulae
1
= h2 (u 2 + 12 + v 2 ),
2

(8.2)

1
= h2 [s1 (u0 )2 + s2 (10 )2 + s3 (v 0 )2 + 2r1 h1 1 u0
2
+2r2 h1 1 v 0 + 2r3 h1 v10 + 22 h2 12 + 32 h2 v 2 ].

(8.3)

and

In (8.2) and (8.3) u(x, t) is the average longitudinal displacement (with some
normalization weight), 1 (x, t) and v(x, t) describe the two thickness-stretch
and thickness-shear branches of longitudinal vibrations, the principal terms
of which can be regarded, in the long-wave range, as the first eigenfunctions
of vibrations of the cross section as a two-dimensional continuum. It is
assumed that these unknown functions are continuous and have continuous
derivatives with respect to x and t. The density of the 1-D kinetic energy
represents the simplest quadratic form of u,
1 and v that is positive definite.
The density of the 1-D strain energy , in contrast to the general quadratic
forms of u, 1 , v and u0 , 10 , v 0 , possesses some special features, namely, i) there
are no terms of the type u2 , 10 u, u0 v, uv 0 , u0 v 0 , u0 10 , and ii) the quadratic
form containing only the derivative with respect to x is positive definite.
The construction of the theory is completed by specifying the constants
s1 , s2 , s3 , r1 , r2 , r3 , 2 , 3 , which depend only on the material constants and
the geometry of the cross section, and indicating the way to restore the 3-D
displacement field of the rod from the functions u, 1 and v. The next three
sections will be devoted to this problem.
Flexural vibrations. We assume, additionally, that the rod has a plane of
symmetry and that vibrations occur in that plane. Then the kinetic and
strain energy of the rod are given by
1
= h2 (u 21 + h2 2 ),
2

(8.4)

8.1. ONE-DIMENSIONAL EQUATIONS

315

and
1
= h2 [sh2 ( 0 )2 + 12 ( + u01 )2 ].
2

(8.5)

Here u1 is the transverse displacement of the rod averaged over the cross
section with some weight and h( + u01 ) describes the first thickness-shear
branch of vibrations. It is also understood that u1 and are continuously
differentiable with respect to x and t. One recognizes immediately in (8.4)
and (8.5) Timoshenkos theory of flexural vibrations of beams proposed in
[57,58]. The variational-asymptotic method can be used here to calculate the
coefficients , s, 1 , as well as to restore the three-dimensional displacement
field of the rod from the functions u1 and .
1-D equations of motion
Longitudinal vibrations. We derive now the equations of high-frequency vibrations for the elastic rod from the variational principle (8.1), (8.2), (8.3).
Calculating the variation of the action functional, we have
2

t1

[(u
u + 1 1 + v
v)
(s1 u0 u0 + s2 10 10

J = h

t0
0

0
0

+ s3 v v + r1 h1 1 u0 + r1 h1 u0 1 + r2 h1 1 v 0 + r2 h1 v 0 1
+ r3 h1 10 v + r3 h1 v10 + 22 h2 1 1 + 32 h2 vv] dx dt.
Performing the integration by parts and remembering that u, 1 and v are
specified at t = t0 , t1 , we obtain
2

t1

J = h

t0

{[
u + (s1 u00 + r1 h1 10 )]u + [1 + (s2 100

r1 h1 u0 r23 h1 v 0 22 h2 1 )]1 + [
v + (s3 v 00
Z t1
+ r23 h1 10 32 h2 v)]v} dx dt h2
[(s1 u0 + r1 h1 1 )u
t0

+(s2 10

x=L
+ r3 h v)1 + (s3 v + r2 h1 1 )v] x=0 dt, (8.6)
1

where r23 = r2 r3 . The Euler equations of (8.6) follow

u = (s1 u00 + r1 h1 10 ),
1 = (s2 00 r1 h1 u0 r23 h1 v 0 2 h2 1 ),
1
00

v = (s3 v + r23 h

10

32 h2 v).

(8.7)

316

CHAPTER 8. ELASTIC RODS

Flexural vibrations. The equations of flexural vibrations for the rod can be
derived in a similar manner and are of the form

u1 = 12 ( 0 + u001 ),
h2 = [sh2 00 12 ( + u01 )].

(8.8)

Boundary-value problems
Longitudinal vibrations. Consider the variant of the boundary conditions for
the system (8.7) that corresponds to the free edges of the rod. Similar to the
classical rod theory we assume that the variations of u, 1 , v are arbitrary at
x = 0, L. Then from (8.6) the following boundary conditions can be obtained
at x = 0, L:
s1 u0 + r1 h1 1 = 0,
s2 10 + r3 h1 v = 0,
s3 v 0 + r2 h1 1 = 0.

(8.9)

If the tractions act on the lateral surface, then the variational principle (8.1)
should be modified by adding the term associated with the work done by the
surface tractions:
Z t1 Z L
J =
( + A) dx dt = 0,
(8.10)
t0

where
A = h(p1 u + p2 1 + p3 v).

(8.11)

The quantities p1 , p2 , p3 depend only on the surface traction ti ; their expressions will be determined later. These quantities should be included in the
equations of longitudinal vibrations

u = (s1 u00 + r1 h1 10 + h1 p1 ),
1 = (s2 00 r1 h1 u0 r23 h1 v 0 2 h2 1 + h1 p2 ),
1
00

v = (s3 v + r23 h

10

32 h2 v

2
1

+ h p3 ).

(8.12)

To complete the formulation of the boundary-value problems we specify also


the intinial conditions at t = t0
u|t0 = u0 ,
u|
t0 = u1 ,

1 |t0 = 10 ,
1 |t0 = 11 ,

v|t0 = v 0 ,
v|
t0 = v 1 .

8.1. ONE-DIMENSIONAL EQUATIONS

317

If the coefficients s1 , s2 , s3 are positive, then the system (8.7) (or (8.12)) is
hyperbolic, and the corresponding initial boundary-value problem is wellposed.
Flexural vibrations. The free-edge boundary conditions read
+ u01 = 0,

0 = 0.

(8.13)

If the tractions act on the lateral surface, then the variational principle (8.10)
should be used instead of (8.1), where
A = h(q1 u1 + hq2 + hq3 u01 ).
The quantities q1 , q2 , q3 depend only on ti and will be found later. The
equations of vibrations then read

u1 = 12 ( 0 + u001 + h1 q1 q30 ),
h2 = [sh2 00 2 ( + u0 ) + q2 ].
1

(8.14)

The functions u1 , and their first derivatives with respect to t should be


specified at t = t0
u1 |t0 = u01 ,
u 1 |t0 = u11 ,

|t0 = 0 ,
t = 1.
|
0

The system (8.8) (or (8.14)) is hyperbolic if s > 0, and the corresponding
initial boundary-value problem is well-posed in that case.
Problems
1. Derive the balance equation of energy for a rod within the 1-D theory
of high-frequency vibrations. Using this equation, prove the uniqueness
of solutions of the boundary-value problems.
2. Find out the conditions which guarantee the hyperbolicity of the equations (8.7) and (8.8).
3. Formulate the variational principle for the eigenvalue problems and
prove the orthogonality of the eigenfunctions with respect to the energy.

318

8.2

CHAPTER 8. ELASTIC RODS

Long-wave asymptotic analysis

3-D variational formulation. We first formulate the problem of free vibrations of the rod that is straight and untwisted in its natural state. Consider
the domain B E specified by
B = {(x1 , x2 , x3 )|(x1 , x2 ) S, x3 x (0, L)},
where {xi } is a cartesian co-ordinate system (the superscript 3 of x3 is usually omitted for short). Concerning the plane figure S we assume that it
is connected (but need not be simply connected), centrally symmetric (if
(x1 , xR2 ) S then (x1 , x2 ) S) and (x1 , x2 ) = (0, 0) is its centroid, so
that S x da = 0. A linear elastic body occupying the domain B in its stressfree undeformed state is called an elastic rod, with S its cross section. The
diameter of S is denoted h; when h  L the rod is said to be thin.
To be specific, we will assume that the lateral surface of the rod is free
from surface tractions. Then the true displacements of the rod w , w are
sought as the stationary points of the three-dimensional action functional
Z

t1

I=

da dx dt,
t0

with the Lagrangian given by


1
1
2

2
) + 2w;
w,x + ( + 2)w,x
= (w 2 + w 2 ) [ (w;
2
2
+ 2w(;) w(;) + (w; + w,x )2 ].
The variations of w , w should vanish at t = t0 , t1 and at the rod ends x = 0, L
which are assumed to be clamped.
The first step of the variational-asymptotic procedure. We analyze the action functional given above in the long-wave range, for which the
parameter h/l is small everywhere inside the rod, with l the characteristic
wavelength in the longitudinal directions. However, we do not assume that
the frequency of vibrations is small. By stretching the transverse co-ordinates
with = x /h we get the Lagrangian which depends on h explicitly
1
1
2

2
= (w 2 + w 2 ) [h2 (w|
) + 2h1 w|
w,x + e2 w,x
2
2
+ 2h2 w(|) w(|) + (h1 w| + w,x )2 ],

(8.15)

8.2. LONG-WAVE ASYMPTOTIC ANALYSIS

319

p
where = / and e = /( + 2). We use the vertical bar preceding
a Greek index to denote the partial differentiation with respect to . The
action functional becomes
Z t1 Z L
2
hi dx dt,
(8.16)
I=h
0

t0

where h.i denotes the integral over S = { |x S}.


In the first step of the variational asymptotic procedure we neglect formally all small terms in (8.15) and arrive at the zero approximation functional
2

t1

h0 i dx dt,

I0 = h

t0

(8.17)

1
1
2
2
0 = (w 2 + w 2 ) [h2 (w|
) + 2h2 w(|) w(|) + h2 w|
].
2
2
Observe that, in contrast to the low-frequency vibrations, the kinetic energy
must already be retained in the first step of the variational-asymptotic procedure since the frequencies are no longer small. The derivatives with respect
to x do not enter the functional
R t(8.17); therefore its stationary points coincide
with those of the functional t01 h0 idt. The latter are linear superpositions
of vibrations, which fall into two series, namely:
w = v(n) f(n) ( ), w = 0,
(F Lk (n)),

w = 0, w = (n) f(n) ( ), (T F L (n)).

(8.18)
(8.19)

The series F Lk contains the branches of flexural-longitudinal vibrations, the


series T F L the branches of torsional-flexural-longitudinal vibrations.1
The subscripts and k denote the thickness-stretch and thickness-shear
branches, respectively. It is understood that the functions v(n) , (n) depend
harmonically on t with frequency . In each branch the frequency and eigenfunction can be determined by solving the following eigenvalue problems
F Lk :

2
2 f(n) = k(n)
f(n) ,

f(n)| = 0 at S,

(8.20)

and
T F L :

2
f(n) ,
( + 1)f(n)|
2 f(n) = (n)

f + 2f((n)|) = 0 at S.

(8.21)

(n)|

The more detailed classification is possible for centrally symmetric cross sections, for
which w and w are decomposed into even and odd functions.

320

CHAPTER 8. ELASTIC RODS

Here = h/cs , 2 is the 2-D Laplace operator, and are the components
Without restricting the generality, one
of the unit normal to the contour S.
can impose on f(n) and f((n) the following normalization conditions
2
hf(n)
i = 1,

hf(n) f(n)
i = 1.

(8.22)

These conditions are chosen so as to simplify the average kinetic energy.


For functions v(n) , (n) independent of x, each of the solutions given above
represents an exact solution of the equations of 3-D elasticity for an infinite
rod and corresponds to the synchronized vibrations of cross sections along
the rod (with the zero longitudinal wave number). The frequencies k(n) and
(n) will be called cut-off frequencies. For vibrations whose amplitude and
frequency vary slowly in the longitudinal directions of the rod, the equations
(8.18)-(8.19) can be regarded as the zero approximations, for which
2
v(n) k(n)
v(n) ,

2
(n) (n)
(n) ,

(8.23)

except those corresponding to the zero frequency (they belong to the kernels of the operators). The displacements of these classical low-frequency
branches of vibrations, in the first step of the variational-asymptotic procedure, turn out to be independent from, or linear functions of, (see
Chapter 4). All the remaining branches correspond to vibrations with frequency cs /h. Since as h 0, the corresponding vibrations
are naturally called high-frequency (or thickness) vibrations. We rewrite the
eigenvalue problems (8.20), (8.21) as the following variational problems
F Lk :

hf| f | i = k2 hf f i,

(8.24)

hf 2 i = 1,
and
T F L :

hf| f|
+ 2f(|) f (|) i = 2 hf f i,

(8.25)

hf f i = 1.
In the following the eigenvalues and the corresponding eigenfunctions are no
longer numbered in order to avoid complicated notations.
The second step of the variational-asymptotic procedure. At this
step we find the next refinement for the displacements of each branch of
vibrations. Consider a branch belonging to the series F Lk . Regarding v as
a given function of x and t satisfying the constraints
v = 0 at x = 0, L,

(8.26)

8.2. LONG-WAVE ASYMPTOTIC ANALYSIS

321

we seek w for that branch. Keeping the principal terms depending on w


and the principal cross terms between w and v in (8.16), we obtain the
functional
Z t1 Z L
2
h1 i dx dt,
(8.27)
I1 = h
t0

1
1
0
2
vf
1 = k2 w2 [h2 (w|
) + 2h1 w|
2
2
+2h2 w(|) w(|) 2h1 w v 0 f| ].
The first and last terms of 1 are obtained after the integration by parts with
respect to t and x, respectively. While doing so, the term w
is replaced by
k2 w since we are interested in the long-wave asymptotics near the cut-off
frequency. The terms that go to the boundaries x = 0, L vanish due to (8.26).
From (8.27) it follows that the displacements of the branch F Lk are given by
w = vf ( ),

w = hv 0 g ( ),

(8.28)

where g ( ) is the solution of the following variational problem

hg|
g|
+ 2g(|) g (|) + f g|
f| g i = k2 hg g i.

(8.29)

Consider now a branch belonging to the series T F L . Regarding as a


given function of x and t satisfying the constraint
= 0 at x = 0, L,

(8.30)

we seek w for that branch. To do this we must retain the principal terms
containing w and the principal cross terms between w and in (8.16). As a
result we obtain the functional (8.27) with the following Lagrangian
1 2 2 1
2

1 =
w (h2 w|
+ 2h1 0 f w| 2h1 0 f|
w).
2
2
Again, the first and last terms in this Lagrangian were obtained by the
integration by parts taking the condition (8.30) into account. Consequently,
the following asymptotic formulae hold true for the displacement distribution
of the branch T F L
w = f ( ),

w = h 0 g( ),

(8.31)

where g( ) is the solution of the following variational problem

hg| g | + f g | f|
gi = 2 hggi.

(8.32)

322

CHAPTER 8. ELASTIC RODS

Average Lagrangian of an individual thickness branch. Before performing the short-wave extrapolation it is first convenient to find the principal
terms of the average Lagrangian of each branch in the long-wave range. We
examine a branch in the series F Lk , whose displacements are given by the
asymptotic formulae (8.28). Let v(x, t) now be an arbitrary function of x
and t. Substituting (8.28) into the 3-D action functional (8.16), retaining
the principal terms and averaging over the cross section, we obtain
= 1 [v 2 + h2 (v 0 )2 hg g i] 1 [h2 v 2 hf| f | i

2
2
0 2
2
(|)

+ (v ) (h(g| ) + 2g(|) g
+ 2f g|
2g f| i + e2 )].
Due to the variational equations (8.24), (8.29) and the normalization condition we have
hf| f | i = k2 ,
2

h(g|
) + 2g(|) g (|) + f g|
f| g i = k2 hg g i = ck2 ,

where c = hg g i. Consequently
= 1 [v 2 + ch2 (v 0 )2 ] 1 [k2 h2 v 2 + (v 0 )2 (ck2 + kk )],

2
2

(8.33)

kk = e2 + hf g|
g f| i.

(8.34)

where

Since v describes harmonic vibrations in the long-wave range near the cut-off
frequency k , for which the estimation (8.23) holds in the first approximation,
the small correction term ch2 (v 0 )2 in (8.33) can be replaced by ck2 h2 (v 0 )2 .
Then the principal terms of the branch F Lk in the average Lagrangian are
given by the following final formula
= 1 v 2 1 [ 2 h2 v 2 + kk (v 0 )2 ].

k
2
2

(8.35)

Should this branch be uncoupled from the other branches, the Euler equation
of (8.35) would be of the form

v = (kk v 00 k2 h2 v).

(8.36)

One can show that (8.36) is asymptotically exact in the long-wave range near
the cut-off frequency k .
Consider now a branch in the series T F L , whose displacement field in
the long-wave range is given by (8.31). We substitute these formulae into

8.2. LONG-WAVE ASYMPTOTIC ANALYSIS

323

the action functional (8.16). After discarding the small terms and averaging
over the cross section we arrive at
= 1 [ 2 + h2 ( 0 )2 hg 2 i] 1 [h2 2 h(f )2 + 2f(|) f (|) i

|
2
2

2
g + 2f g| i + 1)].
2f|
+ ( 0 )2 (hg|
Making use of the variational equations (8.25) for f and (8.32) for g one can
show that
2
) + 2f(|) f (|) i = 2 ,
h(f|
2

hg|
f|
g + f g| i = 2 hg 2 i = 2 d,

where d = hg 2 i. Thus,
= 1 [ 2 + h2 d( 0 )2 ] 1 [h2 2 2 + (k + d2 )( 0 )2 ],

2
2
where

k = hf g| f|
gi + 1.

(8.37)

With the same deliberation as in the previous case one can replace the term
2 2
dh2 ( 0 )2 in this Lagrangian by d
h ( 0 )2 . Thus, we derive the final formula
for the average Lagrangian of the branch T F L
= 1 2 1 [h2 2 2 + k2 ( 0 )2 ].

2
2

(8.38)

Not only the principal terms containing the factor 1 in the kinetic energy
and the factor h2 in the strain energy, but also terms of the next order of
smallness must be retained in the average Lagrangians (8.35) and (8.38), due
to the fact that, at the cut-up frequencies, the sum of the principal terms
turns out to be small. Regarding the branch T F L as uncoupled from the
other branches, we would obtain for the following Euler equation
= (k 00 h2 2 ) = 0,

(8.39)

which is asymptotically exact in the long-wave range near the cut-off frequency of this branch.
Problems
1. Derive the equations and boundary conditions for g from (8.32) and
for g from (8.29).

324

CHAPTER 8. ELASTIC RODS

2. Find out the conditions which guarantee the hyperbolicity of the equations (8.36) and (8.39).

3. Prove the orthogonality between different thickness branches with respect to the energy in the long-wave range, provided the edges of the
rod are clamped.

8.3

Short-wave extrapolation

Longitudinal vibrations. We now consider free longitudinal vibrations


of the rod with arbitrary boundary conditions at its edges (clamped or free
edges). We assume that these vibrations can be regarded, with sufficient
accuracy, as the superposition of the first three branches of the longitudinal vibrations. The branch Lk (0) correspond to low-frequency vibrations,
the other ones to longitudinal thickness vibrations with the lowest cut-off
frequencies, in which the most essential part of the vibrational energy is concentrated. Besides, the necessity of including also the first thickness-shear
branch into the theory is dictated by its strong interaction with the first
thickness-stretch branch, similar to the shell and plate theory. The dynamic
equations contain three unknown functions of x and t: u, 1 , v (the symbols
without the bar are reserved for the functions in the final equations). Despite the fact that the theory involves more unknown functions than in the
classical rod theory, it should be regarded as a first approximation theory
describing asymptotically exactly the vibrations of the rod in the range of
long waves and high frequencies (up to the second cut-off frequency).
Thus, we represent the rod displacements in the form
w = uc + vf + h10 g,
w = 1 f + h
u0 e + h
v 0 g ,

(8.40)

where the functions u, 1 , v are arbitrary functions of x and t. The function


u corresponds to the classical longitudinal branch of the vibrations and describes the mean longitudinal displacement of the rod (with some weight).
The functions 1 and v correspond to the first longitudinal thickness branches
in the series T F L and F Lk . The basis functions f and f are orthonormal
eigenfunctions in variational problems (8.24) and (8.25), and the functions g
and g are determined in terms of f and f as the solutions of problems (8.29)
and (8.32). From the asymptotic analysis of the low-frequency branch of the
1/2 = const
longitudinal vibrations (Chapter 4) it is easy to see that c = |S|
while e is a linear function in , which is a solution of the variational

8.3. SHORT-WAVE EXTRAPOLATION

325

problem
he| e| + 2e(|) e(|) + ce| i = 0,
he i = 0,

(8.41)

he e i = 0.

This problem has the obvious solution


e = c ,

(8.42)

where is Poissons ratio.


We substitute (8.40) into the 3-D action functional (8.16) and integrate
Keeping the principal terms of each branch and the
over the cross section S.
principal cross terms in the average Lagrangian and taking account of the
results in the previous section, we obtain
= 1 (u 2 + 2b1 hu 0 + 2 + 2b2 h 1 u 0 + 2b3 h 1 v 0 + v 2 + 2b4 hv 0 )

1
1
1
2
1
[22 h2 12 + 2a1 h1 1 u0 + 2a2 h1 1 v0 + 2r1 h1 u0 1
2
+ 2r2 h1 1 v0 + 2r3 h1 v10 + 32 h2 v2 + 2a3 h1 v10
+ k1 (
u0 )2 + k2 (10 )2 + k3 (
v 0 )2 ]. (8.43)
Here 2 and 3 are the first eigenvalues and k of the longitudinal vibrations in problems (8.24) and (8.25) (they are numbered analogously to the
case of longitudinal vibrations of plates). The formulae for the remaining
coefficients have the form
b1 = hcgi, b2 = hf e i, b3 = hf g i, b4 = hf gi,


a1 = hf|
e| + 2f(|) e(|) i, a2 = hf|
g| + 2f(|) g (|) i,

a3 = hf| g | i, r1 = hf|
ci, r2 = hf|
f i,

r3 = hf| f i,

(8.44)

k1 = E/ = 2(1 + ).

The coefficients k2 , k3 are given by the formulae (8.37) and (8.34), respectively.
It turns out that the following relationships between the coefficients
b1 = b2 , b3 = b4 , a1 = b2 22 , a2 = b3 22 , a3 = b4 32 ,
r1 = b1 22 = a1 , r2 r3 = b3 (32 22 ) = a3 a2

(8.45)

hold true. This can be proved by using the variational problems (8.24),
(8.25), (8.29), (8.32) and (8.41), without calculating the coefficients. Indeed,

326

CHAPTER 8. ELASTIC RODS

the substitutions f = e for f in (8.25) and g = c for g in (8.32) lead


to
hf| e| + 2f(|) e(|) i = 22 hf e i = 22 b2 ,

(8.46)

hf|
ci = 22 hgci = 22 b1 .

(8.47)

It follows from (8.44), (8.46) and (8.47) that the equalities a1 = 22 b2 and
r1 = 22 b1 are valid. Subtracting (8.47) from (8.46), we obtain the equality

ci = 22 (b2 b1 ).
hf| e| + 2f(|) e(|) + f|

(8.48)

The left-hand side of (8.48) coincides with the variational equation (8.41)
for e , when we substitute therein e = f . Therefore 22 (b2 b1 ) = 0, or
b2 = b1 .
The substitutions f = g in (8.24), f = g in (8.25), g = f in (8.29)
and g = f in (8.32) yield
hf| g | i = 32 hf gi = 32 b4 ,

(8.49)

hf| g|
+ 2f(|) g (|) i = 22 hf g i = 22 b3 ,

hg|
f| + 2g(|) f (|) + f f|
f| f i

(8.50)

= 32 hg f i = 32 b3 ,

(8.51)

hg| f

+ f f

f|
fi

22 hf gi

22 b4 .

(8.52)

From (8.49) and (8.50) follow a3 = 32 b4 and a2 = 22 b3 . Subtracting (8.51)


from (8.50), we get the equality

hf| f f f|
i = b3 (22 32 ).

According to (8.44) this equality means r2 r3 = b3 (32 22 ). Finally, adding


(8.51) and (8.52) together and subtracting (8.49) and (8.50) from the result
we get
(22 32 )(b4 b3 ) = 0.

(8.53)

Since 2 6= 3 , it follows from (8.53) that b3 = b4 . Thus, relations (8.45) are


proved.
It follows from relations (8.45) that the principal cross terms in the average Lagrangian (8.43) form divergent terms (null Lagrangian) that do not
affect the equations of vibrations in the long-wave range. Consequently, a
method of short-wave extrapolation is possible in which all the cross terms in
(8.43) are discarded, which would result in uncoupling between the vibration

8.3. SHORT-WAVE EXTRAPOLATION

327

branches. However, additional analysis shows that such extrapolation leads


to a qualitatively false description of the dispersion curves and integral characteristics of the rod in the short-wave range (see Section 8.5). Therefore,
we will here keep all the cross terms and just try to seek the change of the
unknown functions that would simplify (8.43), i.e., would reduce (8.43) to
an expression without second and higher-order derivatives with respect to x
and t.
Using (8.45) we transform the Lagrangian (8.43) to
= 1 [(u + b2 h 0 )2 + ( 1 + b2 hu 0 + b3 hv 0 )2 + (v + b3 h 0 )2

1
1
2
1
(b22 + b23 )h2 ( 10 )2 b22 h2 (u 0 )2 b23 h2 (v 0 )2 ] [22 h2 (1 + b2 h
u0 + b3 h
v 0 )2
2
v + b3 h10 )2
+ 2r1 h1 u0 1 + 2r2 h1 1 v0 + 2r3 h1 v10 + 32 h2 (
+ (k1 b22 22 )(
u0 )2 + (k2 b23 32 )(10 )2 + (k3 b23 22 )(
v 0 )2 ]. (8.54)
When transforming (8.43) to (8.54) the cross terms h2 u 0 v 0 and u0 v0 are neglected as small compared with other cross terms. In (8.54) the terms in
the kinetic energy 1/2(b22 + b23 )h2 ( 10 )2 and 1/2b23 h2 (v 0 )2 at long waves
v 0 )2 (as was done in
can be replaced by 1/2(b22 + b23 )22 (10 )2 and 1/2b23 (
2 2 0 2
Section 8.2). The term 1/2b2 h (u ) is small at long waves and can be
omitted. The Lagrangian then has the form
= 1 [(u + b2 h 0 )2 + ( 1 + b2 hu 0 + b3 hv 0 )2 + (v + b3 h 0 )2 ]

1
1
2
1
u0 + b3 h
v 0 )2 + 2r1 h1 u0 1 + 2r2 h1 1 v0
[22 h2 (1 + b2 h
2
+ 2r3 h1 v10 + 32 h2 (
v + b3 h10 )2 + (k1 b22 22 )(
u0 )2
+ (k2 b23 32 + (b22 + b23 )22 )(10 )2 + (k3 b23 22 + b23 32 )(
v 0 )2 ]. (8.55)
In order to search for a short-wave extrapolation which does not contain
second and higher derivatives in the Lagrangian let us make the changes of
unknown functions u u, 1 1 and v v, where
u = u + b2 h10 ,
1 = 1 + b2 h
u0 + b3 h
v0,
v = v + b3 h10 .
Keeping the principal terms of u, 1 , v in (8.55), we arrive at
= 1 (u 2 + 12 + v 2 ) 1 [22 h2 12 + 32 h2 v 2 + 2r1 h1 u0 1

2
2
0 2
2r1 b2 (u ) 2r1 b2 1 100 + 2r2 h1 1 v 0 2r2 b3 1 100 2r2 b3 (v 0 )2

328

CHAPTER 8. ELASTIC RODS


+2r3 h1 v10 2r3 b3 (10 )2 2r3 b3 vv 00 + (k1 b22 22 )(
u0 )2
v 0 )2 ].
+(k2 b23 32 + (b22 + b23 )22 )(10 )2 + (k3 b23 22 + b23 32 )(

By adding a null Lagrangian, we can replace terms of the type 1 100 and
vv 00 by (10 )2 and (v 0 )2 , respectively, and derive finally the formulae (8.2)
and (8.3) for and , where the coefficients s1 , s2 , s3 are given by
s1 = k1 b22 22 2r1 b2 ,
s2 = k2 b23 32 + (b22 + b23 )22 + 2r1 b2 + 2r2 b3 2r3 b3 ,
s3 = k3 b23 22 + b23 32 2r2 b3 + 2r3 b3 .
Making use of relations (8.45) again, we obtain
r12
s1 = k1 + 2 ,
2
r2
r2
s2 = k2 12 + 2 23 2 ,
2
3 2
2
r
s3 = k3 2 23 2 .
3 2

(8.56)

When the tractions ti act on the lateral boundary of the rod we must add
the work A done by ti to the action functional. With the displacements from
(8.40), the principal terms of A are found to be given by (8.11), where
Z
Z
Z
1
1
1

tc ds, p2 =
t f ds, p3 =
tf ds.
p1 =
S
S
S
Thus, the formulae of the 1-D theory of high-frequency longitudinal vibrations is justified by the short-wave extrapolation procedure, which consists of
the operations of adding or removing i) terms that are small in the long-wave
range, or ii) null Lagrangians. Besides, by solving the cross section problems
we can calculate all the coefficients of the theory.
Flexural vibrations. Assume that (x1 , x) is the plane of symmetry of the
cross section, and vibrations occurring in that plane can be regarded with
sufficient accuracy as the superposition of the first two branches of the flexural vibrations. The branch F (0) correspond to low-frequency vibrations, the
other one to thickness-shear vibrations with the lowest cut-off frequencies.
The dynamic equations contain two unknown functions of x and t: u1 and
(the symbols without the bar are reserved for the functions in the final
equations). The theory should be regarded as a first approximation theory
describing asymptotically exactly the flexural vibrations of the rod in the

8.3. SHORT-WAVE EXTRAPOLATION

329

range of long waves and high frequencies (up to the cut-off frequency of the
thickness-shear branch).
We represent the displacement field in the form
w = u1 (x, t)c + h2 u001 m ( ),
t)f ( ).
w = h
u01 c1 1 + (x,

(8.57)

The function u1 describes the means displacement (with some weight) of the
rod in the plane (x1 , x), where c are constants chosen so that
1/2 ,
c1 = |S|

c2 = 0.

According to (4.43) the functions m ( ) are given by


1
h i
m ( ) = ( c + c c )(
).
2
|S|
The functions correspond to the first flexural thickness branch in the series
Fk . The basis function f is an odd with respect to eigenfunction of
variational problem (8.24), for which another normalization condition will
be used. In (8.57) we omit the term h0 g , which gives a contribution (0 )2
to the average Lagrangian. This is due to an additional analysis, which shows
that a hyperbolic short-wave extrapolation describing exactly the curvature
of the dispersion curve near the cut-up frequency of the first branch of Fk
does not exist.
Let us calculate the strain field
33 = w,x = h
u001 c1 1 + 0 f,
1
1
23 = w,x + w| = h
y,x + f
| ,
h
h
1
 = w(|) = h
u001 m(|) = h
u001 c1 1 .
h

(8.58)

We substitute (8.57) into the 3-D action functional (8.16) and integrate over
Using the decomposition (4.29) and (8.58) one can see
the cross section S.
that the transverse energy is small compared with other terms in the average
Lagrangian. Keeping in the latter the principal terms, we obtain
= 1 (u 21 + 2 ) 1 [sh2 (

u001 )2 2r1 h
u001 0
2
2
1 2
2

+ 1 2 + 2r2 h
u000
1 ],
h

(8.59)

330

CHAPTER 8. ELASTIC RODS

where the coefficients , 1 , s, r1 , r2 are given by


= hf 2 i, 12 = hf| f | i,
E
E
s = c21 h( 1 )2 i, r1 = hc1 1 f i,

r2 = hm f | i.

(8.60)

It can be shown that, if the next correction term for the classical branch
were taken into account, all the cross terms in (8.59) would form a null
Lagrangian. However, here we shall keep the cross-terms and try to seek
appropriate changes of unknown functions to get rid of terms containing the
higher order derivatives. For this aim let us transform (8.59) to
r1 1 0 2
= 1 (u 2 + 2 ) 1 [sh2 (
u001
)

1
2
2
sh
1
r2
2
+ 12 2 ( + 2 h3 u000
1 ) ].
h
1

(8.61)

2
Small terms of the type (0 )2 and h4 (
u000
1 ) are neglected when passing from
(8.59) to (8.61). We now choose the normalization condition for f such that

r1
hc1 1 f i
= 2 1 2 = 1.
s
c1 h( ) i

(8.62)

Making the changes of unknown functions u1 u1 and in accordance


with
u1 = u1 ,

r2
1
=
u01 + ( + 2 h3 u000
1 ),
h
1

and keeping in the average Lagrangian only the principal terms, we arrive at
the formulae (8.4), (8.5).
Problems
1. Check the validity of (8.56).
2. Prove that the short-wave extrapolations proposed are asymptotically
equivalent to the classical theory of vibrations of rods in the lowfrequency limit.
3. Find the expressions of the work done by the tractions acting on the
lateral boundary of the rod in the case of the flexual vibrations.

8.4. CROSS SECTION PROBLEMS

8.4

331

Cross section problems

Longitudinal vibrations. The coefficients of the theory can be determined


by solving the cross section problems (8.24), (8.25), (8.29), (8.32). They
are rather complicated and can, in general, be solved only with the help of
numerical methods. Exceptional cases are circle and circular ring, for which
analytical solutions are available. Since computations for the circular ring are
rather cumbersome, we consider, for simplicity, only circular cross sections.
Let us introduce the polar co-ordinates %, . Looking for axially symmetric
vibrations we have, obviously, f g 0, while f, f% , g, g% depend only on
%. Therefore, problems (8.24), (8.25), (8.29), (8.32) reduce to boundary-value
problems for ordinary Bessel differential equations and are solved explicitly
in terms of Bessels functions.
We first determine the functions f satisfying the boundary value problem
(8.21). Seeking f in the form f = p| , where p depends only on %, we reduce
(8.21) to the equation

1 d dp
(% ) = 22 p,
% d% d%

2 = 2 e.

This equation yields the following non-singular solution


p = c2 J0 (2 %),

f% = c2 2 J1 (2 %).

(8.63)

From (8.21) the boundary condition for p follows

1 d dp
d2 p
(% ) + 2 2 = 0,
% d% d%
d%

at % = 1/2.

Substitution of (8.63) into (8.64) leads to the equation for 2


4e2 J1 (2 /2) = 2 J0 (2 /2).
The constant c2 is determined from the normalization condition
Z 1/2

f%2 %d% = 1.
hf f i = 2
0

With f% from (8.63) we find that


c2 =

21
,
2 J0 (2 /2)

2e
1 = p 2
.
2 /4 4(1 e2 )

Summing up, we obtain


21
f% =
J1 (2 %).
J0 (2 /2)

(8.64)

332

CHAPTER 8. ELASTIC RODS

We now turn to the function f , which should satisfy problem (8.20). Since
f depends only on %, problem (8.20) in the polar co-ordinates becomes

1 d df
(% ) = 32 f,
% d% d%

df
= 0 at % = 1/2.
d%

It follows from here that


f = c3 J0 (3 %),
where 3 is the lowest positive root of
J1 (3 /2) = 0 3 = 7.6634.
Using the normalization condition (8.22) it is easy to show that

c3 = 2/ J0 (3 /2).
Thus, the function f is found to be
f=

2
J0 (3 %).
J0 (3 /2)

(8.65)

Having found f , we can determine g by solving problem (8.32), which is


nothing but

2 g ( + 1)f|
= 22 g,

(g| + f ) = 0 at S.

= 2 p and rewriting this equation in the polar co-ordinates,


Replacing f|
we obtain

1 d dg
2( + 1)1 2
(% ) + 22 g =
J0 (2 %),
% d% d%
J0 (2 /2)
dg
21 J1 (2 /2)
=
at % = 1/2.
d%
J0 (2 /2)

(8.66)

Problem (8.66) yields the following solution


1
2
1
J0 (2 %)
J0 (2 %)].
g= [
eJ1 (2 /2)
2 J0 (2 /2)
Analogously, substituting f from (8.65) into the problem (8.29), and seeking g in the form g = q| , where q depends only on %, we obtain the following
boundary-value problem for q
1 d dq
2( + 1)e2
(% ) + 32 q =
J0 (3 %),
% d% d%
J0 (3 /2)
d2 q dq
2
e2 2 +
=
at % = 1/2,
d%
% d%

(8.67)

8.4. CROSS SECTION PROBLEMS

333

where 3 = e3 . Problem (8.67) yields the following solution for q


2
4
J0 (3 %),
q=
J ( %) +
2 0 3
2
J0 (3 /2)3
(3 J0 (3 /2) + 43 J1 (3 /2))
and, consequently, g% is equal to
g% =

43
2
J1 (3 %).
J1 (3 %)
2
J0 (3 /2)3
(3 J0 (3 /2) + 43 J1 (3 /2))

Having found all the functions, we now use the formulae (8.34), (8.37)
and (8.44) to determine all the coefficients of the 1-D theory. After long,
but otherwise standard, calculations we arrive at the following expressions
for r1 , r2 , r3
22

r1 = hcf|
i = p 2
,
2 /4 4(1 e2 )
2e2 23

p
r2 = hf|
fi = 2
,
(3 22 ) 22 /4 4(1 e2 )
2 2
p 3 2
r3 = hf f| i = 2
.
(3 22 e2 ) 22 /4 4(1 e2 )

(8.68)

The coefficients k2 and k3 are determined from (8.34), (8.37) and are given
by
42 J0 (2 /2)
22 /(4e2 ) 12
+ 2
,
2
2
2 /4 4(1 e ) (2 /4 4(1 e2 ))J1 (2 /2)
43 J1 (3 /2)
k3 = 1 +
.
3 J1 (3 /2) 32 J0 (3 /2)/4

k2 =

(8.69)

Finally, the coefficients s1 , s2 , s3 are calculated from k2 , k3 and r1 , r2 , r3 according to (8.56). In the following table we present the values of 2 , r1 , r2 ,
r3 , s1 , s2 , s3 as functions of

2
r1
r2
r3
s1
s2
s3

0
5.21
0
0
6.19
2
0.63
1.89

0.5
5.47
0.52
0.17
6.22
2.11
0.62
1.95

0.1
5.77
1.17
0.39
6.25
2.24
0.62
2.03

0.15
6.14
1.98
0.71
6.28
2.40
0.62
2.13

0.2
6.59
3.04
1.17
6.31
2.61
0.61
2.24

0.25
7.17
4.49
1.85
6.34
2.89
0.59
2.38

0.3
7.95
6.63
2.95
6.38
3.29
0.56
2.58

0.35
9.1
10.11
4.89
6.43
3.92
0.48
2.87

0.4
11.04
16.96
8.96
6.48
5.16
0.16
3.42

334

CHAPTER 8. ELASTIC RODS

One can see that for [0, 0.4] the coefficients s1 , s2 , s3 are positive
and the Euler equations (8.7) are hyperbolic. As a consequence, one can
also prove the well-posedness of the corresponding initial boundary-value
problems. For (0.4, 1/2) the coefficient s2 is a quickly oscillating function
of , which changes its sign infinitely many times. The situation is similar
to that of the plate theory and can be explained by the fact that, as
increases, the cut-off frequency 2 of the branch L (0) also increases and
crosses the regions of strong interaction with other high-frequency branches
(Lk (2), Lk (3), and so on). Due to this reason the short-wave extrapolation
proposed above becomes unsatisfactory.
Flexural vibrations. In order to determine the constants , 1 , s of the
theory of flexural vibrations we need to solve the only eigenvalue problem
(8.20) subject to the normalization condition (8.62) for which, in contrast to
the previous case, many solutions are known. The determination of and 1
is thus reduced to an interpretation of these solutions and some additional
computations. Results are given for the cross sections bounded by the circle,
rectangles and ellipses.
Circle. For the circular cross section the equation (8.20), in the polar coordinates %, , admits an antisymmetric solution
f = bJ1 (1 %) cos .
Then the boundary condition f /% = 0 at % = 1/2 leads to the equation
for 1
J10 (1 /2) = 0 = 1 = 3.682.

It is easy to see that c1 = 2/ and s = E/16. The coefficient b is


determined from the normalization condition (8.62) yielding
c1

b=
64

R 1/2
0

%2 J1 (1 %)d%

= 0.8217.

Consequently, the coefficient is equal to


2

= b

1/2

J12 (1 %)%d% 0.0633.

Mindlin and Deresiewicz proposed the correction of the Timoshenko theory,


which, in our notation, gives = 1/16 = 0.0625 (see [41]).

8.4. CROSS SECTION PROBLEMS

335

Rectangles. Consider a rectangular cross section of width a and height 1.


We assume for definiteness that a < 1. Problem (8.20) yields the solution
independent of 2

f = b sin 1 1 , 1 = .
a

For the rectangular cross section c1 = 1/ a and s = Ea2 /12. The coefficient b is determined from the normalization condition (8.62) and is equal
to
2
c1 h( 1 )2 i
=
c
a
.
b= 1
1
h sin 1 1 i
24
The coefficient can now be calculated

1 2
= b2 hsin2 ( 1 )i = ( )2 a2 0.0846a2 .
a
2 24
According to Mindlin-Deresiewiczs correction =
[41]).

1 2
a
12

0.0833a2 (see

Ellipses. Consider now an elliptical cross section with the major axis 1 and
the minor axis a, and let a 1 (see Figure 8.1). A solution of the cross

Figure 8.1: Elliptical cross section.


section problem (8.20) can be obtained by separation of variables in elliptical
co-ordinates. Thus, let
1 = e sinh u sin v,
e=

2 = e cosh u cos v,

1 a2 /2 being the half of eccentricity, (8.20) becomes


2
2f
2f
(
+
) + 12 f = 0.
e2 (cosh 2u cos 2v) u2
v 2

(8.70)

336

CHAPTER 8. ELASTIC RODS

Assuming a solution of the form


f = f1 (u)f2 (v)
and substituting it into (8.70) one obtains, after dividing by f
1 d2 f2 e2 12
1 d2 f1 e2 12
cosh
2u
=

cosh 2v.
+
+
f1 du2
2
f2 dv 2
2
Thus, we see immediately that both sides of this equation should be a constant, which we denote by b. The function f2 , which must be periodic of
period 2, satisfies Mathieus equation
d2 f 2
+ (b 2q cos 2v)f2 = 0,
dv 2

(8.71)

with q = e2 12 /4, while f1 is a solution of modified Mathieus equation


d2 f 1
(b 2q cosh 2u)f1 = 0,
du2

(8.72)

subject to the boundary condition


df1
= 0,
du

at u = tanh1 a.

(8.73)

The solutions of these equations are given in terms of Mathieus and radial

Figure 8.2: Cut-off frequency 1 versus a.


Mathieus functions [1]. Denoting b = b1 (q) the first characteristic value of
(8.71), we have
f2 = se(b1 , q, v),

8.4. CROSS SECTION PROBLEMS

337

where se1 (b1 , q, v) is the odd periodic Mathieu function with the characteristic
value b1 and parameter q. Substituting b = b1 (q) into (8.72), we find its
general solution in the form
f1 (u) = cSe(b1 , q, u),
with Se(b1 , q, u) the radial Mathieu function with the parameter b1 and q.
Combining these formulae and substituting the solution into (8.73), we obtain
the equation determining the cut-off frequency 1 . In Figure 8.2 the graph of
1 versus a is plotted. Then we should use the normalization condition (8.62)
to determine the constant c. Finally, knowing f we calculate according to
the formula (8.60). In Figure 8.3 we plot as a function of a, where the
dashed curve corresponds to Mindlin-Deresiewiczs correction = a2 /16.

Figure 8.3: Coefficient versus a.

Problems

1. Solve the cross section problems for the longitudinal vibrations of rods
with a cross section bounded by a circular ring and determine the
corresponding coefficients.
2. Determine Poissons ratio c for which 2 = 3 (c 0.2833).
3. Show that, as c , the coefficients k2 and k3 approach infinity, but
s2 and s3 remain finite. Find them.
4. Plot the eigenfunctions at the cut-off frequencies for the circular cross
section.

338

8.5

CHAPTER 8. ELASTIC RODS

Dispersion of waves

Longitudinal waves. The elastic rod, as waveguide, demonstrates very


complicated behaviour in the high-frequency range, which we want to study
in this section. Within the 1-D theory of rods the longitudinal waves are
described by equations (8.7), which, in terms of the dimensionless variables
=

tcs
,
h

x
,
h

(8.74)

may be written as
u| = s1 u00 + r1 10 ,
1| = s2 100 r1 u0 r23 v 0 22 1 ,
v| = s3 v 00 + r23 10 32 v.

(8.75)

Note that the prime is used here to denote the derivative with respect to .
We seek solutions of (8.75) in the form
(u, 1 , v) = (a, b, c)ei( ) ,

(8.76)

where is the dimensionless wave number and the dimensionless frequency.


The equations (8.75) and solutions (8.76) are quite similar to (7.67) and
(7.69), so that the dispersion relation for (8.76) can immediately be written
down
(s1 2 2 )(s2 2 2 + 22 )(s3 2 2 + 32 )
2 2
(s3 2 2 + 32 )r12 2 (s1 2 2 )r23
= 0.

(8.77)

Analogously, the asymptotic formulae of the dispersion curves in the longwave range (small ) for the three branches Lk (0), L (0) and Lk (1) read
2 = (s1 r12 /22 )2 + O(4 ) = 2(1 + )2 + O(4 ),
2
2 = 22 + (s2 + r12 /22 r23
/(32 22 ))2 + O(4 )
= 22 + k2 2 + O(4 ),
2
2 = 32 + (s2 + r23
/(32 22 ))2 + O(4 )
= 32 + k3 2 + O(4 ),

(8.78)

where the coefficients k2 , k3 are given by (8.69).


We first examine the asymptotic accuracy of the formulae (8.78). The
formula (8.78)1 has already been compared with the exact asymptotics in
Section 4.4. Now we want to derive the exact asymptotics of the thickness branches near the cut-off frequencies from the Pochhammer dispersion

8.5. DISPERSION OF WAVES

339

relation (4.85). Since p1 and p2 are real near the cut-off frequencies, the
Pochhammer equation takes the form
4p1 (p22 + 2 )J1 (p1 /2)J1 (p2 /2) (p22 2 )2 J0 (p1 /2)J1 (p2 /2)
42 p1 p2 J1 (p1 /2)J0 (p2 /2) = 0, (8.79)

where p1 = e2 2 2 , p2 = 2 2 , J0 and J1 being Bessels functions.


The cut-off frequencies are determined from (8.79) by putting = 0. We
then have p1 = e, p2 = , and equation (8.79) becomes
4e3 J1 (e/2)J1 (/2) 4 J0 (e/2)J1 (/2) = 0.
Thus, the first two cut-off frequencies are identical with 2 , 3 . Let us now
compute the curvature of the dispersion curve near the cut-off frequency 3 .
Introduce the following notations
p

2 = x, 2 = y, p1 = e2 y x, p2 = y x,
(x, y) = 4p1 yJ1 (p1 /2)J1 (p2 /2) (y 2x)2 J0 (p1 /2)J1 (p2 /2)
4xp1 p2 J1 (p1 /2)J0 (p2 /2).
Expand the function (x, y) near the point (x, y) = (0, 32 ) to obtain
(x, y) = ax + b(y 32 ) + o(x, y 32 ) = 0,
where



a=
,
x (0, 2 )
3

(8.80)



b=
.
y (0, 2 )
3

Rather long, but otherwise elementary, calculations show that


32
J0 (3 /2) 53 J1 (3 /2)],
4
2
b = J0 (3 /2)3 [3 J1 (3 /2) 3 J0 (3 /2)].
4

a = J0 (3 /2)3 [

(8.81)

Solving (8.80) with respect to y


a
y = 32 x + O(x2 ),
b
and taking (8.81) into account, we arrive at the formula (8.78)3 .
Analogously, expanding (x, y) near the point (x, y) = (0, 22 ), we have
(x, y) = ax + b(y 22 ) + o(x, y 22 ),

(8.82)

340

CHAPTER 8. ELASTIC RODS

where, now,


a=
,
x (0, 2 )



b=
.
y (0, 2 )

After some calculations we find that


2
1
a = J0 (2 /2)22 [42 J0 (2 /2) + J1 (2 /2)(12 22 )],
4
4e
1
b = J0 (2 /2)J1 (2 /2)22 (22 /4 4(1 e2 )].
4

(8.83)

Solving in (8.82) y through x and substituting for a, b their expressions, we


obtain the asymptotic formula (8.78)2 . Thus, the asymptotic exactness of
the 2-D dispersion relation in the long-wave range is proved.

Figure 8.4: Dispersion curves of L-waves: a) 1-D theory of high frequency


vibrations: dashed line and b) 3-D theory: solid line.
In the short-wave range ( ) all the dispersion curves according to
(8.77) approach the asymptotes
=

s2 ,

s3 ,

s1 ,

confirming the hyperbolicity of the 1-D equations (8.75). The dispersion


curves according to the 1- and 3-D theories are plotted in Figure 8.4 for =
0.3317, where the projections of the complex branch onto the real (Re, )
and imaginary (Im, ) planes are marked with r and i, respectively. One can
see that, in addition to the asymptotic exactness in the long-wave range, the

8.5. DISPERSION OF WAVES

341

1-D theory of high-frequency vibrations of rods describes satisfactorily the


first complex branch of the dispersion curves, which gives reason to expect a
good prediction of the exponentially decayed boundary layer by this theory.
Flexural waves. We now turn to the dispersion of flexural waves propagating in an infinite rod of circular cross section. By introducing the dimensionless variables (8.74) we rewrite the equations of flexural motion (8.8) in
the form
u1| = 12 (0 + u001 ),
| = s00 12 ( + u01 ),

(8.84)

where = h. We seek solutions of these equations in form of the harmonic


waves
= (a, b)ei( ) ,
(u1 , )
with the wave number and the frequency. Because of the similarity
between (8.84) and (7.62) we can easily obtain the dispersion relation for the
flexural waves in the rod
(12 2 2 )(s2 + 12 2 ) 2 14 2 = 0.

(8.85)

In the long-wave range (  1) the asymptotic analysis of (8.85) yields the


following formula
2 = s4 + O(6 ),
for the branch F (0), and
2 = 12 + (12 +

s 2
) + O(4 ),

for the branch Fk (0).


In the short-wave range ( ) both the dispersion curves according
to (8.85) approach the asymptotes
r

s
= 1 , =
,

in contrast to the classical 1-D theory. This property is characteristic for


hyperbolic systems. The dispersion curves according to the 1- and 3-D theories are plotted in Figure 8.5 for = 0.31, where a qualitative agreement
between the 1-D theory of high-frequency vibrations and 3-D elasticity is
observed even in the short-wave range.

342

CHAPTER 8. ELASTIC RODS

Figure 8.5: Dispersion curves of F-waves: a) 1-D theory of high frequency


vibrations: dashed line and b) 3-D theory: solid line.
Problems
1. Check the validity of the formula (8.81).
2. Do the same for (8.83).
3. Study the behaviour of the dispersion curves for the longitudinal waves
when Poisson ratio approaches c 0.2833 from below and above
(c is Poissons ratio for which 2 = 3 ).

8.6

Frequency spectra

Longitudinal vibrations. Consider the free longitudinal vibrations of a


rod of circular cross section, which are governed by the equations of motion
(8.75). In the case of harmonic vibrations the solution takes the form
u = u()ei ,

1 = 1 ()ei ,

v = v()ei ,

where is the dimensionless frequency. Then we reduce (8.75) to


s1 u00 + r1 10 + 2 u = 0,
s2 100 r1 u0 r23 v0 + (2 22 )1 = 0,
s3 v00 + r23 0 + (2 2 )
v = 0.
1

(8.86)

Similarly, the traction-free boundary conditions at = d = L/2h become


s1 u0 + r1 1 = 0,
s2 10 + r3 v = 0,
(8.87)
0

s3 v + r2 1 = 0.

8.6. FREQUENCY SPECTRA

343

We introduce three scalar potentials i such that


u = 01 + 02 + 03 ,
1 = a1 1 + a2 2 + a3 3 ,
v =

b1 01

b2 02

(8.88)

b3 03 .

It turns out that u, 1 , v given by (8.88) are the solutions of equations (8.86)
if the potentials i satisfy the following equations
00i + 2i i = 0,

i = 1, 2, 3,

(8.89)

where i are the roots of the dispersion relation (8.77) and ai , bi are expressed
through i by
s1 2i 2
,
r1
r23 ai
bi =
.
2
s3 i + 32 2

ai =

The general solution of (8.89) has the form


i = ci cos i + di sin i .
We express u, 1 , v through i according to the last equation
u =

3
X

ci i sin i +

i=1

1 =

3
X

di i cos i ,

i=1

ai ci cos i +

i=1

v =

3
X

3
X

ai di sin i ,

(8.90)

i=1

3
X

bi ci i sin i +

i=1

3
X

bi di i cos i .

i=1

We substitute these formulae into the boundary conditions (8.87). Since


cos and sin are even and odd functions of , respectively, we obtain
two uncoupled system of homogeneous linear equations for ci and di
3
X

Cij cj = 0,

i = 1, 2, 3,

Dij dj = 0,

i = 1, 2, 3,

j=1

and
3
X
j=1

344

CHAPTER 8. ELASTIC RODS

where
C1j = cos j d,
C2j = j (s2 aj + r3 bj ) sin j d,
C3j = (s3 2j bj + r2 aj ) cos j d,
and
D1j = sin j d,
D2j = j (s2 aj + r3 bj ) cos j d,
D3j = (s3 2j bj + r2 aj ) sin j d.
Thus, the free longitudinal vibrations of the rod fall into two classes, namely,
symmetric vibrations whose eigenfunctions contain ci and antisymmetric ones
whose eigenfunctions contain di . We further restrict ourselves to the symmetric vibrations, the eigenfrequencies of which are determined as the roots
of the equation
det Cij = 0.

(8.91)

Figure 8.6: Frequency spectra of the longitudinal vibrations of rods: a) 1-D


theory of high frequency vibrations: solid line and b) 1-D classical theory:
dashed line.
Figure 8.6 shows the first seven roots of this equation as functions of
the dimensionless length of rods d for = 0.29. The frequencies computed

8.6. FREQUENCY SPECTRA

345

by using the classical theory of longitudinal vibrations are also shown (the
dashed hyperbolas). They are given by
p
2(1 + )
1
(n ).
(8.92)
n =
d
2
One can see that the frequencies n (d) computed in accordance with (8.91)
and (8.92) differ litle for . 3, and the hyperbolas (8.92) become asymptotes for the curves obtained from (8.91) as 0 (d ). For & 3
fairly strong divergence between the curves obtained from (8.91) and (8.92)
is observed.
Edge modes. We want to show that the zone of frequencies, where the
typical terrace-like structure of the spectra begins to develop, contains the socalled edge resonance frequency. Since the latter cannot be described within
the classical rod theory, the frequency spectra as obtained by the theories
of low and high frequency vibrations diverge strongly in its neighborhood.
Indeed, take the point (d, ) = (2, 5.6427) lying on the third terrace in Figure
8.6 and compute the amplitudes of the displacements u, 1 , v according to
(8.90). The results are presented in Figure 8.7. One can see that u, 1 , v

Figure 8.7: Distributions of u, 1 and v at the frequency of edge resonance


(d = 2, e = 5.6427).
become very large as approaches the free ends of the rod. If we restore the
3-D distribution of the displacement field and draw the deformed edge of the
rod, a picture similar to that in 7.16 will be obtained. This shows clearly
that e = 5.6427 correspond to the edge resonance frequency.
The other interesting feature of the frequency spectra shown in Figure
8.6 is that, for 6= 0, the edge mode continuously changes to other modes

346

CHAPTER 8. ELASTIC RODS

of vibrations as d changes, i.e., there is no pure edge mode for finite rods,
whose frequency is a continuous function of d that does not tend to zero as
d . In order to understand this mode switching, let us first analyze
the degenerate case = 0. From (8.68) it follows then that r1 = r2 = 0;
consequently, the eigenvalue problem breaks into the two following problems
s1 u00 + 2 u = 0,

u0 = 0 at = d,

(8.93)

and
s2 100 + r3 v0 + (2 22 )1 = 0,
v = 0,
s3 v00 r3 10 + (2 32 )
0
0
s2 1 + r3 v = 0, v = 0, at = d.

(8.94)

Thus, for = 0 the classical and thickness branches of vibrations are uncoupled. The symmetric modes of vibrations of (8.93) possess the eigenfrequencies given by (8.92). The symmetric solution of (8.94) have the form
1 = c1 cos 1 + c2 cos 2 ,
v = b1 c1 1 sin 1 b2 c2 2 sin 2 ,

(8.95)

where 21 , 22 are the roots of the quadratic equation


(s2 2i 2 + 22 )(s3 2i 2 + 32 ) r3 2i = 0,

(8.96)

and bi are the coefficients given by


bi = r3 /(s3 2i 2 + 32 ),

i = 1, 2.

Substituting the general solution (8.95) into the free edge boundary conditions and equating the determinant to zero, we obtain the frequency equation
in the form
tan(1 d)
(s2 + r3 b2 )1 b1
=
.
tan(2 d)
(s2 + r3 b1 )2 b2

(8.97)

We now analyze the asymptotics of the solution of (8.97) as d in the


interval of frequencies < ? , where the equation (8.96) has two complexconjugate roots 21 , 22 . It is easy to see that in the limit d the lefthand side of equation (8.97) tan(1 d)/ tan(2 d) 1 so that the curve (d)
approaches the horizontal line = , where is the root of the equation
(s2 + r3 b1 )2 b2 = (s2 + r3 b2 )1 b1 .

(8.98)

8.6. FREQUENCY SPECTRA

347

Figure 8.8: Small pertubation of the spectra for small positive .


Equation (8.98) is nothing but the equation determining the frequency of
the edge mode in a semi-infinite rod with = 0 (compare with the similar
equation (7.97)). Solving (8.98) with = 0, 2 = 5.208, 3 = 7.663 and
s2 = 0.625, s3 = 1.886, r3 = 6.194, we find 4.666, which differ from
that value T 4.73 computed by 3-D elasticity [18] by about 1.3%. Thus,
for = 0 there exists the pure edge mode, whose frequency depends continuously on d and approaches 4.666 as d . Figure 8.8 shows
the frequency spectra of the rods in the neighborhood of at large d (for
= 0). There is a countable set of intersection points of the hyperbolas
(8.92) with the horizontal line = corresponding to the multiple frequencies in the degenerate case = 0. Now, let Poissons ratio be a small
number differing from zero. Then r1 , r2 6= 0 and the small coupling between
v occurs. According to the theory of perturbations [45] the multiplicity
u, ,
of the frequencies should disappear. Obviously, the solid lines in Figure 8.8
change to the dashed lines which are already similar to those in Figure 8.6.
One can determine the edge resonance frequency versus Poissons ratio
in exactly the same manner as in the case of the semi-infinite trip. Namely,
given the incident wave propagating toward the free edge of the rod, one
seeks the amplitudes of the reflected waves, two of which having complex
conjugate wave numbers corresponding to the localized motion near the free
edge. At a certain frequency e the phase difference between the reflected
and the incident waves vanishes, and, simultaneously, the amplitudes of the
waves with the complex wave numbers reach their maximum, which can be
interpreted as the edge resonance vibration. In Figure 8.9 the dependence of
the edge resonance frequency on Poissons ratio (0 0.3) is shown,
where small circles correspond to the edge resonance frequency computed by
3-D elasticity [18].

348

CHAPTER 8. ELASTIC RODS

Figure 8.9: Edge resonance frequencies in semi-infinite rods versus Poissons ratio .
Flexural vibrations. The flexural vibrations of the rod in the plane 1 ,
are governed by the equations (8.84). For free harmonic vibrations of the
type
u1 = u1 ()ei ,

= ()e
,

these equations reduce to


12 (0 + u001 ) + 2 u1 = 0,
s00 12 ( + u01 ) + 2 = 0.

(8.99)

The general solution of (8.99) reads


u1 =

2
X

bi cos i +

2
X

i=1

2
X

ci sin i ,

i=1

ai bi i sin i +

i=1

2
X

ai ci i cos i ,

i=1

where bi , ci are arbitrary constants, 2i are the roots of the quadratic equations
(8.85), and ai are given by
ai =

2 12 2i
,
12 2i

i = 1, 2.

We substitute this general solution into the free edge boundary conditions at
= d
+ u01 = 0,

0 = 0.

8.6. FREQUENCY SPECTRA

349

Figure 8.10: Frequency spectra of the flexural vibrations of rods.


Due to evenness and oddity of the cosine and sine functions, respectively,
two uncoupled systems of homogeneous linear equations for bi and ci are
obtained
2
X

Bij bj = 0,

i = 1, 2,

Cij cj = 0,

i = 1, 2,

j=1

and
2
X
j=1

where
1
sin j d,
j
B2j = (2 12 2j ) cos j d,
B1j =

and
1
cos j d,
j
C2j = (2 12 2j ) sin j d,

C1j =

The free flexural vibrations of the rod also fall into two classes, namely, symmetric vibrations whose eigenfunctions contain bi and antisymmetric ones

350

CHAPTER 8. ELASTIC RODS

whose eigenfunctions contain ci . We further restrict ourselves to the symmetric vibrations, the eigenfrequencies of which are determined as the roots
of the equation
det Bij = 0.
In Figure 8.10 we plot the roots of this equation as functions of the
dimensionless length of rods d for = 0.29. In contrast to the previous case,
the terrace-like spectra begins to develop at the frequencies larger than the
cut-off one 1 ,2 and no edge mode is observed in this case.
Problems
1. Study the frequency spectra of the antisymmetric longitudinal vibrations of rods.
2. Study the frequency spectra of the antisymmetric flexural vibrations of
rods.
3. Derive the equation determining the edge resonance frequency of the
semi-infinite rod. In which case does an exponentially decayed solution
exist? Find it.

For = 0.29 the cut-off frequency 1 3.682.

Chapter 9
Piezoelectric shells
9.1

Two-dimensional equations

Variational principle. Consider a piezoelectric shell with a middle surface


S and a thickness h as described in Section 5.1. For simplicity we assume
that the material of the shell is transversely isotropic and homogeneous (the
piezoceramic shell with thickness polarization). Let the face surfaces of the
shell be covered by electrodes, and the time-dependent voltage, 0 (t), be
specified on them. The problem is to construct a two-dimensional theory of
piezoceramic shells that enables one to determine their resonant vibrations
in a wide range of frequencies. Taking the variational approach as the basis,
we postulate that the vibrations of the piezoceramic shell should occur in
accordance with the following variational principle
Z

t1

Z
( ) da dt = 0,

J =
t0

(9.1)

where and are the 2-D kinetic energy and electric enthalpy density, respectively. The functional (9.1) is called, as before, the 2-D action functional.
In the theory of low-frequency vibrations of fully electroded piezoelectric
shells the 2-D action functional, as it was shown in Section 5.1, depends only
on the three functions u (x , t) and u(x , t) describing the mean displacements of the shell. It is natural to assume that, as the frequency increases,
some internal degrees of freedom corresponding to branches of the shell thickness vibrations will be excited and become more and more involved, so that
they should be included as unknown functions in the 2-D kinetic energy and
electric enthalpy densities. Within this approach the most important steps
consist in finding i) the list of external and internal degrees of freedom, and
ii) the 2-D kinetic energy and electric enthalpy densities depending on these
351

352

CHAPTER 9. PIEZOELECTRIC SHELLS

degrees of freedom. Since we are concerned here with the resonant vibrations,
whose frequencies are among mechanical eigenfrequencies under short-circuit
conditions, we can set 0 = 0. The essence of the proposed theory of highfrequency vibrations of piezoceramic shells is then expressed by the following
formulae
h
(9.2)
= (u 2 + u 2 + h2 2 + 2 + v 2 ),
2
and
h
h2
2
G[s1 (A )2 + 2A A + (N ( )2 + 2 ) + s2 ,
2
12
2

+ s3 (v;
) + 2v (;) v(;) + 2r1 h1 A + 2r2 h1 v;
+ 2r3 h1 v ,

+ f1 ( + )( + ) + f2 h2 2 + f3 h2 v v ], (9.3)
where = u, + b u , and the short notation X2 = a X X is used to
denote the squared magnitude of the 2-D vector X. In (9.2)-(9.3) the 2D densities of the kinetic energy and electric enthalpy depend on the three
functions external degrees of freedom u , u and on five additional functions
, and v, which represent the most essential internal degrees of freedom
within the range of frequencies of interest. It is assumed that these unknown
functions are continuous and have continuous derivatives with respect to x
and t. The measures of extension and bending are expressed by
A = u(;) b u,
= (;) + b( $) ,
1
$ = (u, u, ).
2
The derivation of (9.2) and (9.3) as well as the determination of the
constants will be given in the next two sections.
Boundary-value problems. Up to the coefficients, the 2-D kinetic energy and electric enthalpy densities (9.2) and (9.3) are of the same form as
the corresponding densities (7.2) and (7.3) for elastic shells. Therefore the
equations of motion for piezoceramic shells can immediately be derived

h
u = t
; q b ,

h
u = q;
+ t b ,
h3 = q m ,

(9.4)

h =
h
v =

Gh(s2 r1 h
r23 h1 v;
f2 h2 ),
2

Gh[s3 v;
+ 2(v(;) ); + r23 h1 , f3
h v ],
2

9.2. LONG-WAVE ASYMPTOTIC ANALYSIS

353

where r23 = r2 r3 and


1
t = n + (b m b m ).
2
The constitutive equations read
n = Gh(s1 A a + 2A + r1 h1 a ),
h3
m = G (N a + 2 ),
12

q = Ghf1 ( + u, + b u ).
If the shell edge is free, the following boundary conditions should be satisfied
t = 0, q = 0, m = 0,
(s2 ; + r3 h1 v ) = 0,

(s3 v;
a

+ 2v

(;)

(9.5)

+ r2 h a ) = 0.

To complete the formulation of the boundary-value problems we specify also


the initial conditions at t = t0
u |t0 = u0 , u|t0 = u0 , |t0 = 0 , |t0 = 0 , v |t0 = v0 ,
t = 1 , v |t = v 1 .
u |t0 = u1 , u|
t0 = u1 , |t0 = 1 , |
0
0

Problems
1. Derive the equations of motion and the boundary conditions for a plate
(with b = 0) in terms of u , , v and u, .
2. Find out the conditions for which the system of equations (9.4) is hyperbolic.
3. Formulate the variational principle for the eigenvalue problem and
prove the orthogonality of the eigenfunctions with respect to the energy.

9.2

Long-wave asymptotic analysis

Three-dimensional action functional. Let us consider a domain B E


of the form (5.1). A linear piezoceramic body occupying the domain B in its
stress-free undeformed state is called a piezoceramic shell, the surface S its
middle surface, and h its thickness. The shell is said to be thin if h is much

354

CHAPTER 9. PIEZOELECTRIC SHELLS

smaller than the characteristic sizes as well as the characteristic radius of


curvature of the middle surface. Let the piezoceramic shell be polarized along
the thickness, and let its face surfaces be covered by electrodes. According
to the variational principle (2.45) and (2.47) the true displacement field w
i
and the electric potential correspond to stationary points of the action
functional
Z Z Z
h/2

t1

(T W ) dx da dt

I=
t0

(9.6)

h/2

under the short-circuit conditions


|x=h/2 = 0,
where = 1 2Hx + Kx2 and da denotes the area element. In (9.6) T is
the kinetic energy density
1
1
T = w i w i = (a w w + w 2 ),
2
2
with w , w the projections of the displacement vector onto the tangential and
normal directions to the middle surface
w = ti wi ,

w = ni wi .

The electric enthalpy density W is the quadratic form of the strain ab and the
electric field Ea , which, for the piezoceramic shell with thickness polarization,
is given by
W =

G
[(g  )2 + 2g g   + 4g 3 3 + c233 + 2g  33
2
2dg  E3 2e33 E3 4f g 3 E g E E E32 ],

where g are the contravariant components of the metrics given by (3.36),


and where the electroelastic constants, in the abbreviated indicial notation,
are equal to
E
cE
cE
11 2c66
G=
=
, = 66 ,
G
G
E
cE
c
e
31
c = 33 , = 13 , d =
,
G
G
G
e33
e15
S11
S33
e=
, f=
, =
, =
.
G
G
G
G

cE
55 ,

(9.7)

9.2. LONG-WAVE ASYMPTOTIC ANALYSIS

355

The components of the strain tensor are expressed in terms of the displacements in accordance with (3.39), while the components of the electric field
read
E = , ,

E3 = ,x .

Let us stretch the co-ordinate x as follows


x = x/h.
The aim of this change is to express explicitly the dependence of the functional I on the small parameter h. The electric enthalpy now becomes
W =

G
[(g  )2 + 2g g   + 4g 3 3 + c233
2
+ 2g  33 + 2dh1 g  | + 2eh1 33 | + 4f g 3 ,
g , , h2 (| )2 ], (9.8)

where the components of the strain tensor are given by (3.40). The displacements w , w and the potential are the stationary points of the functional
(9.6) with W from (9.8) under the constraints
|=1/2 = 0.

(9.9)

As before, the variations of w , w should vanish at t0 , t1 . Concerning the


boundary conditions at the shell edge, we first consider the clamped edge for
which
w = 0,

w = 0 at S (1/2, 1/2).

The first step of the variational-asymptotic procedure. We study


the variational problem of finding the extremals of the functional (9.6) in the
high-frequency long-wave range with the help of the variational-asymptotic
method. Assuming the smallness of the parameters h/R and h/l everywhere
in S, and discarding formally all small terms in (9.6) we arrive at the zero
approximation functional
Z

t1

Z Z

I0 = h
t0

1/2

1
1
G
{ w 2 + a w w [ch2 (w| )2
2
2
1/2 2

+ h2 a w| w| + 2eh2 w| | h2 (| )2 ]} d da dt. (9.10)

356

CHAPTER 9. PIEZOELECTRIC SHELLS

The potential should satisfy the constraints (9.9). The stationary points
of the functional (9.10) fall into four classes corresponding to four series of
vibrations. The series of thickness-stretch vibrations are characterized by
w = q(),

= (),

w = 0,

(9.11)

where q(), () correspond to the odd or even solutions of the variational


problem
hcq| q| + eq| | + e| q| | | 2 qqi = 0,
(1/2) = 0.

(9.12)

Here and below the angle brackets h.i denote the integration over within
the interval [1/2, 1/2]. The series with the odd solutions is denoted by L
(longitudinal thickness-stretch vibrations), that with the even solutions by
F (flexural thickness-stretch vibrations). The thickness-shear vibrations are
characterized by
w = = 0, w = v p(),
(9.13)
where p() are odd or even solutions of the variational problem
hp| p| k2 ppi = 0.

(9.14)

The series with the odd solutions is denoted by Fk , that with the even solutions by Lk . The eigenvalues and k run through a countable set of values; however, indices are attached neither to them nor to the corresponding
eigenfunctions in order to avoid complicated notations. Since the solutions
of the variational problems (9.12) and (9.14) are determined uniquely up to
a constant factor, the following normalization conditions can be imposed
hq 2 i = hp2 i = 1.

(9.15)

They are chosen so as to simplify later the average kinetic energy. Further,
the functions , v depend harmonically on t with frequency determined
from the appropriate values of or k according to
=

cs
h

or k =

k cs
,
h

(9.16)

p
where cs = G/. For functions , v independent of x , each of the solutions given above represents an exact solution of the equations of 3-D piezoelectricity for an infinite plate and corresponds to the synchronized vibrations
of transverse fibres along the plate (with the zero longitudinal wave number). The frequencies (9.16) will be called cut-off frequencies. For vibrations

9.2. LONG-WAVE ASYMPTOTIC ANALYSIS

357

whose amplitude and frequency vary slowly in the longitudinal directions of


the plates and shells, the equations (9.11) and (9.13) can be regarded as the
zero approximations.
The second step of the variational-asymptotic procedure. We find
the next refinement for the displacements of branches of the thickness-stretch
vibrations (the series L or F ). Considering belonging to one branch of
these series as a given function of x and t satisfying the condition
= 0 at S,

(9.17)

we seek w for that branch. Keeping the principal terms depending on w


and the principal cross terms in (9.6), we obtain the functional
Z t1 Z Z 1/2
1 d da dt,
(9.18)
I1 = h
t0

1/2

with
1 2 2 G 2
1 =
w [h (w| )2 + 2h1 w| , q 2h1 w , q|
2
2
1
2dh w , | + 2f h1 w| , q].
Integration by parts was performed and terms that go to the boundary vanish
due to (9.17). Thus, the functional does not depend on the derivatives of
w with respect to x and t, and the last two enter it as parameters. The
extremal of (9.18) has the form
w = h, s(),

(9.19)

where s() is the solution of the variational problem


hs| s| + qs| q| s d| s + f s| 2 ssi = 0.

(9.20)

Next, let us seek the corrections to w and


w = q() + w? ,

= () + ? .

Here w is considered fixed and defined by (9.19). Without limiting the


generality, the following constraint can be imposed on w?
hw? qi = 0.
It corresponds to the assumption that u = hwq()i. We discard small terms
containing w? and small cross terms. It turns out that the cross terms of the

358

CHAPTER 9. PIEZOELECTRIC SHELLS

order h2 vanish due to the variational equation (9.12), and the functional
(9.6) takes the form (9.18) with a Lagrangian given by the formula
G
1 2 ? 2
2
(w ) 2
Hhqw? [ch2 (w|? )2 + 2eh2 w|? ?|
1 =
2
2
h2 (?| )2 + 4Hh1 (cq| w|? q| w? qw|?
d| w? dq?| eq| ?| e| w|? + | ?| )].
Its extremal has the form
w? = hHz(),

? = hH(),

where z(), () is the solution of the variational problem


hcz| z| + e| z| + ez| | | | 2cq| z|
2q| z 2qz| 2dq| 2d| z 2eq| |
2e| z| + 2| | 2 zz 22 qzi = 0 (9.21)
under the constraints
hzqi = 0,

(1/2) = 0.

(9.22)

Summarizing, we have the following distribution of the displacements and


electric potential over the thickness in the series of thickness-stretch vibrations (in the first approximation)
w = q() + hHz(), w = h, s(),
= () + hH().

(9.23)

We now turn to the series of thickness-shear vibrations. Fixing v , we


seek w and in the second step for one branch of these series (in the first
step w = = 0). Keeping the principal terms of w and and the principal
cross terms, we obtain the functional (9.18) with
1
G

p| w + ch2 (w| )2 + 2h1 v;


pw|
1 = k2 w2 [2h1 v;
2
2

+ 2dh1 v;
p| + 2eh2 w| | 2f h1 v;
p| h2 (| )2 ].
Integration by parts was performed taking into account that v = 0 at the
boundary. The extremal is easily found to have the form

w = hv;
r(),

= hv;
(),

9.2. LONG-WAVE ASYMPTOTIC ANALYSIS

359

where r(), () is the solution of the variational problem


hp| r + cr| r| + pr| + dp| + e| r|
+ er| | f p| | | k2 rri = 0

(9.24)

under the constraints


(1/2) = 0.

(9.25)

We seek the refined term to w in the form


w = v p() + w? ,
where w? satisfy the following constraints
hw? pi = 0.
Again, after discarding small terms containing w? and small cross terms in
the functional (9.6), we arrive at the functional (9.18) with the following
Lagrangian
1
G
?
)2
1 = k2 (w? )2 2k2 Hhv pw? [h2 (w|
2
2
?
?
4Ha h1 v p| w|
+ 2b h1 v pw|
+ 2b h1 v p| w? ].
The extremal should have the form
w? = hy (v , ),
where y is the solution of the variational equation
hy| y| 2Ha v p| y| + b v py|
+ b v p| y k2 y y + 2k2 Hv py i = 0. (9.26)
Thus, for the series of thickness-shear vibrations we have
w = v p() + hy (v , ),

w = hv;
r(), = hv;
().

(9.27)

Average Lagrangians of individual branches of vibrations. We represent the displacements w, w and the electric potential in the form of the
infinite series of branches given above, where , v are now arbitrary functions of x and t. After substituting these series into the action functional

360

CHAPTER 9. PIEZOELECTRIC SHELLS

(9.6) and integrating over the thickness we neglect those small terms of the
order h/R, h/l compared with 1. It turns out that the thickness branches are
orthogonal relative to the action functional in the long-wave range, provided
the shell edge is clamped. Therefore the average functional has the form
Z

t1

J =h
t0

da dt,

(9.28)

decomposes into the series of average Lawhere the average Lagrangian


grangians of low-frequency and thickness branches.1
We examine a branch in the series of thickness-stretch vibrations, whose
displacements and electric potential are given by the asymptotic formulae
(9.23). Substituting (9.23) into the 3-D action functional (9.6), retaining the
principal terms and averaging over the cross section, we obtain
= ( 2 + l2 h2 a , , + l4 h2 2 )

2
G 2 2 2
(h + l1 a , , + l3 2 ),
2

(9.29)

where the coefficients are of the form


l1 = h(s| + q)2 2sq| 2ds| + 2f (s| + q) 2 i,
l2 = hs2 i,
l3 = 4H 2 + 4(2H 2 K) + hcK 2 (q| )2 4cH 2 q| z|
+cH 2 (z| )2 8(H 2 K)qq| 4H 2 qz|
4H 2 zq| + 4(2H 2 K)qq| 8d(H 2 K)q|
4dH 2 z| 4dH 2 q| + 4d(2H 2 K)q|
+2eK 2 q| | 4eH 2 q| | 4eH 2 z| |
+2eH 2 z| | K 2 (| )2 + 4H 2 | | H 2 (| )2 i,
l4 = hK 2 q 2 + H 2 z 2 4H 2 qzi.

(9.30)

Within the first-order approximation we can further simplify the average


Lagrangian. Indeed, integrating the terms l2 h2 a , , and l4 h2 2 over t
2
by parts and replacing by
, the average Lagrangian for the thicknessstretch branch becomes
= 2 G [(h2 2 + i ) 2 + k a , , ].

2
2
1

The average Lagrangian of the low-frequency branches is given in Section 5.2.

(9.31)

9.2. LONG-WAVE ASYMPTOTIC ANALYSIS

361

Making use of the variational equations (9.20) and (9.21), we can simplify
the expressions for k and i leading to
k = hqs| + q 2 sq| ds| + f s| + 2f q 2 i,
i = 4H 2 + 4(2H 2 K) + hcK 2 (q| )2 + 4Kqq|
+4dKq| + 2eK 2 q| | K 2 (| )2 cH 2 (z| )2
+H 2 (| )2 2 K 2 q 2 + 2 H 2 z 2 i.

(9.32)

We turn to a branch in the series of thickness-shear vibrations, whose


displacements and electric potential are given by the asymptotic formulae
(9.27). Substituting (9.27) into the 3-D action functional (9.6), retaining the
principal terms and averaging over the thickness, we get
= [v 2 + l2 h2 (v )2 + l4 h2 v v ] G [h2 2 v 2

;
k
2
2
2
+l1 (v;
) + 2v (;) v(;) + l3 v v ],

(9.33)

where the coefficients are given by


l1 = hp2 2rp| + c(r| )2 + 2pr| + 2dp|
+2es| | 2f p| (| )2 i,
l2 = hr2 i,
l3 v v = h2(Hb Ka )v v 2 (p| )2 4Ha v p| y|
+a y| y| + 4(b Ha )b v v pp|

(9.34)

+2b v py| + c v v p + 2b v p| y
4(b Ha )b v v 2 (p| )2 2c v v pp| + c v v 2 (p| )2 i,
l4 v v = ha K 2 p2 v v 4a H 2 v py + a y y i.
2
) and l4 h2 v v over t by parts and replacIntegrating the terms l2 h2 (v ,
ing v by k2 v , we reduce the average Lagrangian for the thickness-shear
branch to
= v 2 G [(h2 2 a + i )v v

k
k
2
2
(9.35)
(;)
2
+2v
v(;) + kk (v; ) ].

Making use of the variational equations (9.24) and (9.26), we can simplify
the expressions for kk and i
k yielding
kk = hp2 rp| + pr| + dp| f p| i,
2
2

2
i
k v v = h(Ka (p| ) 2Ka pp| + c p

k2 a K 2 p2 )v v a y| y| + k2 a y y i.

(9.36)

362

CHAPTER 9. PIEZOELECTRIC SHELLS

By varying the action functional (9.28) with the average Lagrangians

from (9.29) and (9.33) we derive the equations of long-wave thickness


vibrations, which are similar to those for elastic shells.
Problems
1. Determine the cut-off frequencies and the distributions of displacements
and electric potential from the variational problems (9.12) and (9.14)
for the series L and Lk .
2. Check the formulae (9.30) and (9.32).
3. Check the formulae (9.34) and (9.36).

9.3

Short-wave extrapolation

Displacements and electric field. Let us consider free vibrations of the


piezoceramic shell under short-circuit conditions, with arbitrary boundary
conditions at its edge (clamped or free edge). We assume that the vibrations we are going to describe can be regarded with sufficient accuracy as
the superposition of the branches F (0), Fk (0), Lk (0), L (0) and Lk (1). The
branches F (0) and Lk (0) correspond to low-frequency vibrations, the other
ones to thickness vibrations with the lowest frequencies. The reasons for such
a choice are similar to those utilized in the theory of high-frequency vibrations
for elastic shells. The dynamic equations contain eight unknown functions of
v (the symbols without
the longitudinal co-ordinates and time: u, u , , ,
the bar are reserved for the functions in the final equations).
Thus, we represent the displacements and the electric potential of the
piezoceramic shell in the form

w = u + hA m() + h2 n() + a()


+ h
v;
r(),
w = u h + f () + v d () + h, s(),

(9.37)

= h2 () + ()
+ h
v;
().

In (9.37) the functions u, u describe the low-frequency branches which have


been studied in details in Chapter 5.2 The 2-D measures of extension and
2

Strictly speaking, terms of the type h3 ; l() should be included for the consequent
account of all principal cross terms in the average Lagrangian. However, one can show
that the answer does not depend on this, and such terms are dropped here for short.

9.3. SHORT-WAVE EXTRAPOLATION

363

bending are expressed through u , u as follows


A = u(;) b u,
) ,
= (;) + b( $
1
= u, + b u , $
= (
u, u, ).
2
The functions m(), n(), () satisfy the variational equations (cf. the results of Chapter 5)
hcm| m| + m| i = 0, hmi = 0,
hcn| n| + en| | + e| n| | | n| d| i = 0,
(1/2) = 0, hni = 0.
(9.38)
The functions describe the thickness-shear branch Fk (0) which is strongly
coupled with the branch F (0). An additional analysis shows that a hyperbolic short-wave extrapolation taking the interaction between these branches
into account and describing exactly the curvature of the dispersion curve near
the cut-up frequency of Fk (0) does not exist. Therefore, we drop the cor

rection terms of the type h;


r1 () in w and h;
1 () in . The associate

functions f () are given by


f () = g() + hj (),

j () = y ,

(9.39)

where g() is the first odd solution of the variational equation3


hg| g| 12 ggi = 0,

h12gi = 1,

(9.40)

while y is the solution of the problem (9.26), in which p should be replaced


v describe the thickness branches L (0) and
by g. Finally, the functions ,
Lk (1), respectively, where a(), () and d () are given by
a() = q() + hHz(),
d ()

p()

() = () + hH(),

hy (),

y ()
v = y .

The functions q, and z, should be determined from the variational problems (9.12) and (9.21), respectively, while the functions p, y are the solutions
of the corresponding problems (9.14) and (9.26) derived for these branches
in the previous section.
3

We choose for g another normalization condition to simplify the subsequent changes


of unknown functions.

364

CHAPTER 9. PIEZOELECTRIC SHELLS

Short-wave extrapolation of the average Lagrangian. We substitute


the formulae (9.37) into the action functional (9.6) with the electric enthalpy
from (9.8) and integrate over the thickness. Discarding small terms in the
asymptotic sense and using the results of the previous section, we obtain the
average functional (9.28) with
= 1 (u 2 + 2 + u 2 + 2b1 hu ; + 2 + 2b2 h A + 2b3 h v + v 2

2
1
v
+2b4 hv ; ) G[22 h2 2 +32 h2 v2 +2(a1 +r1 )h1 A +2(a2 +r2 )h1
;
2
2
2
) + 2
v (;) v(;)
+ k3 (
v;
+ 2(a3 + r3 )h1 v , + P (A )2 + 2A A + k2 ,
+

h2

(N (
)2 + 2
) + 2hc1 ;
+ 2hc2 (;) + h2 12 2
12


2
+ 2hc3 ; + i
v ]. (9.41)
1 + i2 + i3 v

In this Lagrangian the coefficients 1 , 2 , 3 are the eigenvalues of the variational problems for g, q, and p, respectively; the coefficients P , N should
have the form
P = + hm| i,

N = 12hn| + d| i,

(9.42)

while k2 , k3 , i
1 , i2 , i3 are calculated from the corresponding eigenfunctions
according to the formulae (9.32) and (9.36) for k , i and kk , i
k . The remaining coefficients are given by

= hg 2 i, b1 = hsi, b2 = hqmi, b3 = hqri, b4 = hpsi,


c1 = hgi + hgn| + dg| i, c2 = 2hgi,
c3 = hg| ni, a1 = hcq| m| + e| m| i,
a2 = hcq| r| + eq| | + e| r| | | i,
a3 = hp| s| i, r1 = hq| + d| i
r2 = hq| p + d| pi, r3 = hp| q + f p| i.

(9.43)

It turns out that these coefficients satisfy the following noticeable relations
c1 = N /12, c2 = 2/12, b1 = b2 , b3 = b4 ,
a1 = r1 = b1 22 , a2 = b3 22 , a3 = b3 32 ,
r3 r2 = a2 a3 = b3 (22 32 ).

(9.44)

These relations are the direct consequences of the variational equations (9.12),
(9.14), (9.20), (9.24) and (9.38). For example, setting in (9.12) q = m, =

9.3. SHORT-WAVE EXTRAPOLATION

365

0, we obtain the identity a1 = b2 22 . Setting now in (9.20) s = 1 we get


r1 = b1 22 . Substracting the second equation from the first one and taking
into account (9.38) we can prove that (b2 b1 )22 = 0, which means that
b2 = b1 . Now, setting in (9.12) q = r, = and in (9.24) r = q, = ,
respectively, we obtain a2 = b3 22 and a3 = b3 32 . The difference of these
equations leads to r2 r3 = b3 (32 22 ). Taking into account the equations
(9.14) and (9.20) we can prove that r2 r3 = b4 (32 22 ). Since 3 6= 2 , we
obtain b3 = b4 .
in the form
These relations enable one to rewrite
2
= 1 [u 2 + 2 + (u + b1 h , )2 + ( + b1 hA + b3 hv ;
)

2
1
+ (v + b3 h , )2 ] G[22 h2 ( + b1 hA + b3 h
v )2
2
v + 2r3 h1 v ,
+ 32 h2 (
v + b3 h, )2 + 2r1 h1 A + 2r2 h1
;
2 2

2
2
+ (P 2 b )(A ) + 2A A + (k2 + b + b2 2 b2 2 )2
1

1 2

3 2

3 3

N
2
(h
;
)
12
2
c3
+
(h
(;) )(h
(;) ) + h2 12 ( + 2 h3 ; )2
12
1

2

+ i + i2 + i v v ]. (9.45)
2
+ (k3 + b23 32 b23 22 )(
v;
) + 2
v (;) v(;) +

2
While transforming (9.41) to (9.45) terms of the type h2 (A )2 , A v;
, (;
)
4
2
and h (
; ) are neglected as small compared with the remaining term in
2
the long-wave range. Besides, the small correction terms of the type h2 ,
2 2
2 2
and h (v ; ) are replaced by the asymptotically equivalent terms G2 ,
2
and G32 (v;
) , respectively.
We make the following changes of unknown functions
c3
u = u, = + h1 ( + 2 h3 ; ),
1

u = u + b1 h, , = + b1 hA + b3 h
v;
,
(9.46)

v = v + b3 h, .

Retaining the principal terms of u, u , , , v and extrapolating the average


Lagrangian to short waves, we arrive finally at the formulae (9.2) and (9.3),
where
s1 = P + r12 /22 ,
s2 = k2 r12 /22 + (r3 r2 )2 /(32 22 ),
s3 = k3 (r3 r2 )2 /(32 22 ),

(9.47)

366

CHAPTER 9. PIEZOELECTRIC SHELLS


f1 = 12 a + h2 i
1 ,
f2 = 22 a + h2 i2 ,
f3 = 32 a + h2 i
3 .

Solutions of the thickness problems. We now solve the variational problems for the branches of vibrations under consideration, and then calculate
the coefficients of the 2-D theory.
Low-frequency branches. We determine the functions m, n, . From the first
variational problem of (9.38) it is easily seen that m = (/c). From the
second problem we can derive the following equations
cn| + e| = 0,
en| | d = 0,

(9.48)

and the boundary conditions at = 1/2


1
cn| + e| = 0,
2

(1/2) = 0.

(9.49)

Problem (9.48) and (9.49) yields the following solution


d 1 2
( 1/4),
2
ed 1
1
n = ( + ) ( 2 1/12),
c
2
=

(9.50)

where = + e2 /c, d = d e/c.


Branch Fk (0). The Euler equation of the problem (9.14) reads
g| + 12 g = 0,

g| = 0 at = 1/2.

Its odd solution is given by


g() = A sin 1 ,

1 = .

(9.51)

The normalization condition h12gi = 1 gives for A the value A = 2 /24.


We now determine the functions y from the variational problem (9.26) with
p replaced by g. The Euler equations of this problem are
y| + 12 y = 2H g| ,
and the boundary conditions read
y| + b g = 0 at = 1/2.

9.3. SHORT-WAVE EXTRAPOLATION

367

With g = A sin 1 we obtain the solution


1
y = A(H sin 1 + b cos 1 ),
1
where b = b + H .
Branch L (0). The Euler equations of the problem (9.12) have the form
cq| + e| + 22 q = 0,
eq| | = 0,

(9.52)

while the boundary conditions at = 1/2 read


cq| + e| = 0,

= 0.

(9.53)

The odd solution of (9.52) and (9.53) satisfying the normalization condition
hq 2 i = 1 is given by
p
q = Q sin 2 , Q = 2/(1 sin 2 /(2 )),
e
= Q(sin 2 2 sin(2 /2)),
(9.54)

where 2 = c1 , c = c + e2 /, and 2 is the smallest positive root of the


equation
2 /2 =

e2
tan(2 /2).

(9.55)

Introducing x = 2 /2 and 2 = e2 /
c we can rewrite (9.55) in the form
x = 2 tan x. Its smallest positive root depends only on 2 .
We turn to the function s(). It follows from (9.20) that
s| + 22 s + (1 + )q| + (d + f )| = 0,
s| + q = 0 at = 1/2.
Substituting for q and their expressions (9.54) and solving this problem,
we obtain
s = S1 cos 2 + S2 + S3 cos 2 ,
where
S1 = Q1 /(2 ( 2 1)), 1 = 1 + + (d + f )e/,
S2 = 2Q(1 1 ) sin(2 /2)/22 ,
1 2 sin(2 /2)
S3 = Q(1 2
)
.
1 2 sin(2 /2)

368

CHAPTER 9. PIEZOELECTRIC SHELLS

Consider now the functions z(), (). Varying the functional (9.21), we
obtain the system of differential equations
cz| + e| + 22 z = 2cq| 2d| + 2e| ,
ez| + | = 2dq| + 2eq| 2| ,
and boundary conditions at = 1/2
cz| + e| = 2q,

= 0.

The solution of this problem is given by


z = Q sin 2 + Z cos 2 +

4de
2
Q sin
,
2
2
2

e
d Q
= Q sin(2 ) + 2
cos 2

2
e
e
2
+ Z cos 2 2 Q 2 sin
,

2
where
2ed e2
Q
2
1
(c
2)Q + cot
,
c2

2
2
2d Q
e
2
=(
+ Z) cos
.
2
2

Z=

(9.56)

Branch Lk (1). We first determine p(). According to (9.12) we have


p| + 32 p = 0,

p| = 0 at = 1/2.

The first even solution of this equation satisfying the condition hp2 i = 1 is
given by

p = 2 cos 2, 3 = 2.
(9.57)
Knowing p, we next find out r() and () by solving the equations
cr| + e| + 32 r + (1 + )p| = 0,
er| | + (d + f )p| = 0,
and the boundary conditions at = 1/2
cr| + e| + p = 0,

= 0,

9.3. SHORT-WAVE EXTRAPOLATION

369

which can be derived from (9.24). As a result we find that


r = R1 sin 2 + R2 sin 2,
e
= 1 + 2 sin 2 + R2 sin 2,

where

2 2 2
/( 1),
R1 = 1
2
2 2
R2 =
[1 + 2 1 /(1 2 )]/( cos 2 sin ),
2
e
1 = 2 R2 sin ,

2 e
d+f
( 1 2 /( 2 1) +
).
2 =
2

Finally, the functions y can be determined in exactly the same manner as


y . With p from (9.57) we have
y =

2(H v cos 3

1
b v sin 3 ),
3

(9.58)

Calculations of the coefficients. Having found all the functions, we are


now able to calculate the coefficients of the functional (9.1)-(9.3). Substituting (9.50) into (9.42) we obtain
P = + hm| i = 2 /c,
N = + 12hn| d| i = P d2 /
.
With g from (9.51) we find that
1 2
= ( )2 .
2 24
We now calculate r1 , r2 , r3 :
r1 = hq| + d| i = 2Q sin(2 /2),
2 2 2
r2 = hq| p + d| pi = ( + de/) 2 2 2 2 2 Q sin(2 /2),
3 2

2
r3 = hp| (q + f )i = 2Q sin(2 /2)[24 2(1 + 4 ) 2 3 2 2 ],
3 2

370

CHAPTER 9. PIEZOELECTRIC SHELLS

where 4 = f e/. Omitting the long calculations of the coefficients k2 , k3 ,

i
1 , and i3 , we present their final formulae
k2 = 1 + Q2 [2 sin 2 + 3 sin2 (2 /2) cot(2 /2) + 4 sin2 (2 /2)],

2 2
1 2
k3 = + 3 + 2 1
+ 2 2R2 sin [1 +
],
1
1 2
2

i
+ 6Hb ],
1 = [3(H K)a
2

i
+ 6Hb ,
3 = 3(H K)a

where
2
+ 24 5 ,
1 2
1
4 5
2
]
+2
,
2 = [2 1 1 2
2
1 ( 1)
2
3 = 2[1 1 2 /( 2 1)]2 /2 ,
4 5
4 = 4(1 1 )(1 1 + )/22 8 2 2 5 /3,
2
2
2 2
3 = (d + f ) /, 5 = e / .
1 = 1 + 12

We do not present here the formula for i2 which cannot be reduced to a compact form. Note that this coefficient can be calculated numerically using the
formulae (9.58) and (9.56). Finally, the coefficients s1 , s2 , s3 are calculated
in accordance with (9.47).
The table below lists numerical values of the coefficients for shells made
of piezoceramics, whose experimental data are given in Appendix A3:
Material
PZT-4
PZT-5
BaTiO3

2
6.83
7.38
5.81

r1
8.62
10.6
4.39

r2
2.44
3.18
1.54

r3
-6.55
-6.57
-5.74

s1
2.76
3.22
1.39

s2
0.87
0.83
0.74

s3
0.95
1.05
0.54

1.19
1.07
0.97

Problems
1. Show that for the special case of isotropic elastic shells with
= = , = 1, c = + 2, d = e = f = = = 0,
the distributions of displacements over the thickness reduce to those in
Section 7.2.
2. Show that for isotropic elastic shells the coefficients of the 2-D theory
coincide with those given in Section 7.3.

9.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

9.4

371

Frequency spectra of circular plates

Derivation of the frequency equation. When the frequency of the voltage increases approaching the first cut-off frequency of the piezoceramic plate,
the equation (5.74) describing its low-frequency vibrations can no longer be
applied. We now investigate the frequency spectra using the theory of highfrequency vibrations of piezoceramic plates. For the longitudinal vibrations
of the plate the equations of motion (9.4) become

u = G[(s1 + )u; + 2 u + r1 h1 ; ],
= G(s2 2 r1 u r23 h1 v 2 h2 ),
;

v = G[(s3 +

)v;

+ v + r23 h ; 32 h2 v ].
2

Introducing the dimensionless variables


s
x
t G

, =
,
=
h
h
we rewrite these equations in the form
u| = (s1 + )u| + 2 u + r1 | ,

| = s2 2 r1 u| r23 v|
22 ,

v| = (s3 +

)v|

+ v + r23 |

(9.59)
32 v .

Here and below the vertical bar preceding Greek indices denotes the derivatives with respect to the dimensionless co-ordinates and time .
For harmonic vibrations we seek the solution in the form
u = u ( )ei ,

)ei ,
= (

v = v ( )ei ,

where is the dimensionless frequency. Substituting this into (9.59) and


eliminating the common factor ei , we reduce (9.59) to
(s1 + )
u| + 2 u + r1 | + 2 u = 0,
s2 2 r1 u r23 v + (2 2 ) = 0,
|

(s3 +

)
v|

+ v + r23 | + (
2

2
2
3 )
v

(9.60)

= 0.

Analogously, the traction-free boundary conditions (9.5) at the plate edge


read
= 0,
s1 u| + 2
u(|) + r1
s2 | + r3 v = 0,

= 0.
s3 v|
+ 2
v (|) + r2 h1

(9.61)

372

CHAPTER 9. PIEZOELECTRIC SHELLS

Making use of Helmholtzs decomposition theorem one can easily show


that u , and v are expressed in terms of three scalar potentials i , i =
1, 2, 3, and two scalar potentials and $ according to
u = 1| + 2| + 3| + .
. | ,
= a1 1 + a2 2 + a3 3 ,
v = b1 1| + b2 2| + b3 3| +

(9.62)

.
. $| ,

where ai , bi are constants. In a similar manner as in Section 7.5 we can show


that the equations (9.60) will be satisfied if i , and $ satisfy Helmholtzs
equations
2 i + 2i i = 0, i = 1, 2, 3,
2 + 2 = 0,
2 $ + (2 32 )$ = 0,
where 2i , i = 1, 2, 3, are the roots of the equation
(S1 2 2 )(S2 2 2 + 22 )(S3 2 2 + 32 )
2 2
(S3 2 2 + 32 )r12 2 (S1 2 2 )r23
= 0,

(9.63)

and the constants ai , bi are chosen as follows


S1 2i 2
,
r1
r23 ai
bi =
.
2
S3 i + 32 2

ai =

(9.64)

In equations (9.63) and (9.64) S1 = s1 + 2, S2 = s2 , and S3 = s3 + 2.


For axially symmetric vibrations the potentials and $ are set equal to
zero. The remaining potentials i depend only on the radial co-ordinate %
and are given in terms of Bessels function
i = ci J0 (i %),

i = 1, 2, 3.

Substitution of this in (9.62) yields


u% = c1 1 J1 (1 %) c2 2 J1 (2 %) c3 3 J1 (3 %),
= c1 a1 J0 (1 %) + c2 a2 J0 (2 %) + c3 a3 J0 (3 %),
v% = c1 b1 1 J1 (1 %) c2 b2 2 J1 (2 %) c3 b3 3 J1 (3 %).

(9.65)

9.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

373

Referring now the plate to the polar co-ordinates, we write the boundary
conditions (9.61) at % = r/h = r for axially symmetric solutions in the form
u%
d
u% + (S1 2) + r1 = 0,
d%
r
d
S2 + r3 v% = 0,
d%
d
v%
S3 v% + (S3 2) + r2 = 0.
d%
r

S1

Substituting (9.65) into these boundary conditions, we obtain the system of


three homogeneous linear equations for three unknown coefficients c1 , c2 , c3
3
X

Cij cj = 0,

i = 1, 2, 3,

(9.66)

j=1

where
i
J1 (i r) 2 J0 (i r),
r
C2i = (S2 ai + r3 bi )i J1 (i r),
i
C3i = (r2 ai S3 bi 2i )J0 (i r) + 2bi J1 (i r),
r
C1i = 2

i = 1, 2, 3,
i = 1, 2, 3,
i = 1, 2, 3.

A nontrivial solution exists if the determinant in (9.66) vanishes. Thus, the


computation of eigenfrequencies reduces to the solution of the determinantal
equation
det Cij = 0.

(9.67)

Calculation of the spectra. We illustrate the calculation of the spectra


with the example of the piezoceramic plate of barium titanate BaTiO3 , whose
electroelastic constants sE , d, T are given in Appendix A3. The coefficients
of the 2-D equations are taken from the table at the end of the previous section. The determinant in (9.67) is evaluated by exactly the same numerical
program as applied to the frequency equation (7.88). The computation is
simplified by the fact that det C is an even analytic function of i , the roots
of the dispersion equation (9.63). The dispersion curves, expressing the dependence of i on for the different ranges of , are plotted in Figure 9.1.
For < ? one of the 2i is real and the other two are complex conjugate.
If we denote by 23 the complex conjugate of 22 , then the terms of the third

374

CHAPTER 9. PIEZOELECTRIC SHELLS

Figure 9.1: Dispersion curves for barium titanate plate.


column of the determinant in (7.88) are the complex conjugates of the terms
of the second column; i.e.,
Cj3 = Cj2 = Rj2 iIj2 ,
where Rj2 and Ij2 are the real and imaginary parts of Cj2 , respectively.
Hence, adding and substracting the two columns one obtains


C11 R12 I12


det Cij = 4i C21 R22 I22 ,
C31 R32 I32
i.e., the characteristic determinant is purely imaginary.
For (? , 2 ) and > 3 all the three 2i are real and positive, and
the determinant in (9.67) is real. Finally, for lying between the two cut-off
frequencies 2 = 5.809 and 3 = 2, (2 , 3 ), one of the root, say, 21 , is
real but negative, and 1 becomes imaginary: 1 = ik1 . The two remaining
2 and 3 are real and positive. The components Ci1 will alter their form
according to
k1
I1 (k1 r) 2 I0 (k1 r),
r
= (S2 a1 + r3 b1 )k1 I1 (k1 r),

C11 = 2
C21

k1
I1 (k1 r).
r
The frequencies are plotted against the ratio radius to thickness r in Figure
9.2. As r increases all the frequencies diminish and approach the resonance
C31 = (r2 a1 S3 b1 21 )I0 (k1 r) 2b1

9.4. FREQUENCY SPECTRA OF CIRCULAR PLATES

375

Figure 9.2: Frequencies of axisymmetric longitudinal vibrations of barium


titanate plates with free edges as functions of r/h.
frequencies computed by the equation (5.78) of the classical theory of piezoceramic plates. As the frequencies increase, the typical terrace-like structure of
the high-frequency spectrum begins to develop at a frequency, whose asymptotic value, as r , is e 4.37. This frequency is lower than the first
cut-off frequency of the thickness-stretch mode given by = 2 .
Longitudinal axisymmetric vibrations of circular plates made of barium
titanate were studied experimentally by Shaw [53]. He observed the coupling
of the classical branch with other branches whose amplitudes were large
near the boundary of the plate and decreased rapidly toward the centre.
The coupling occurred at the frequency near the terrace-like structure of
the spectrum similar to that in Figure 9.2.4 Computations of u = u% ,
and v = v% for r = 2.9, e = 4.371 (the point lying on the third terrace)
show that the displacements do conform with Shaws observations (see Figure
9.3). Thus, the limiting frequency e 4.37 predicted by the 2-D theory is
lower than the experimental value 4.41 by about 1%. The next terrace-like
structure occurs at the frequency = 5.45, which is again close to Shaws
experimental data.
Problems
4

Shaws experimental value of e was 4.41.

376

CHAPTER 9. PIEZOELECTRIC SHELLS

Figure 9.3: Distributions of u, and v at the frequency of edge resonance


(
r = 2.9, e = 4.371).
1. Study the frequency spectra of circular plates made of PZT-5. Does
the terrace-like structure exist?
2. Compute the functions u, and v for 5.45 near the second terracelike structure of the spectra.
3. Determine the edge resonance frequency of the piezoceramic semiinfinite strip. Compare with that of the piezoceramic circular plate.

9.5

Frequency spectra of cylindrical shells

2-D equations of motion. In this section we study the high-frequency


spectra of piezoceramic closed circular cylindrical shells of the thickness h, the
radius R and the length 2L, the latter being regarded as a parameter which
may vary in some range. Since the 2-D equations of motion for piezoceramic
shells, up to the constants, are identical in form with those for elastic shells,
the result of Section 7.6 can be used with some minor changes. Introducing
the quantities
s
x
h
t G
=
, =
, = h , h? = ,
h
h
R

9.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

377

we try to determine the p


high-frequency spectra in terms of the dimensionless frequency = h/ G/. We write the equations of motion in the
dimensionless form
u1| = [s1 (u1|1 + u2|2 + h? u) + 2u1|1 + r1 ]|1 + (u1|2 + u2|1 )|2
1
h?
h?
+ [ (1|2 + 2|1 ) (u2|1 u1|2 )]|2 ,
12
2
4
1
h?
h?
u2| = (u1|2 + u2|1 )|1 [ (1|2 + 2|1 ) (u2|1 u1|2 )]|1
12
2
4
+[s1 (u1|1 + u2|2 + h? u) + 2(u2|2 + h? u) + r1 ]|2 + h? f12 (2 + u|2 h? u2 ),
u| = f11 (1 + u|1 )|1 + f12 (2 + u|2 h? u2 )|2
h? [s1 (u1|1 + u2|2 + h? u) + 2(u2|2 + h? u) + r1 ],
1
1| = f11 (1 + u|1 ) [N (1|1 + 2|2 ) 21|1 ]|1
12
1
h?
[ (1|2 + 2|1 ) (u2|1 u1|2 )]|2 ,
(9.68)
6 2
4
1
h?
2| = f12 (2 + u|2 h? u2 ) [ (1|2 + 2|1 ) (u2|1 u1|2 )]|1
6 2
4
1
[N (1|1 + 2|2 ) 22|2 ]|2
12
| = s2 (|11 + |22 ) r1 (u1|1 + u2|2 + h? u)
r23 (v1|1 + v2|2 ) f2
v1| = (s3 + )(v1|11 + v2|21 ) + (v1|11 + v1|22 ) + r23 |1 f31 v1 ,
v2| = (s3 + )(v1|12 + v2|22 ) + (v2|11 + v2|22 ) + r23 |2 f32 v2 ,
where
f11 = f111 = 12 + h2? i11
f12 = f122 = 12 + h2? i22
1 ,
1 ,
11
2
2 11
22
2
2 22
f31 = f3 = 3 + h? i1 , f32 = f3 = 3 + h? i1 .
The boundary conditions at the free edges 1 = L/h read
s1 (u1|1 + u2|2 + h? u) + 2u1|1 + r1 = 0,
u1|2 + u2|1 = 0,
1 + u|1 = 0,
N (1|1 + 2|2 ) + 21|1 = 0,
1
h?
(1|2 + 2|1 ) + (u2|1 u1|2 ) = 0,
2
4

(9.69)
(9.70)

378

CHAPTER 9. PIEZOELECTRIC SHELLS


s2 |1 + r3 v1 = 0,
s3 (v1|1 + v2|2 ) + 2v1|1 + r2 = 0,
v1|2 + v2|1 = 0.

Note that the number of the boundary conditions is equal to eight.


High-frequency approximation. At the first glance, the boundary-value
problem posed above seems to be rather complicated due to the coupling
between the unknown functions. However, one can rewrite this problem in
the operator notation as
Lu = 0,
Mu = 0 at 1 = L/h,
where u correspond to the generalized displacements
uT = (u1 , u2 , u, 1 , 2 , , v1 , v2 ),
and L and M are matrix differential operators. It is easy to see that
L = L0 + h? L1 ,
M = M0 + h? M1 ,
where L0 , M0 are the operators according to the plate theory and L1 , M1 are
correction operators.
Thus, if we admit the error h? in determining the eigenfrequencies , the
terms of the order h? in (9.68) and (9.69) can be neglected in the first step.
Seeking the axisymmetric solution with u2 = 2 = v2 = 0 and with the
remaining functions depending only on 1 , the two uncoupled systems are
obtained, namely, the equations
u1| = S1 u1|11 + r1 |1 ,
| = S2 |11 r1 u1|1 r23 v1|1 22 ,
v1| = S3 v1|11 + r23 |1 32 v1 ,
and boundary conditions
S1 u1|1 + r1 = 0,
S2 |1 + r3 v1 = 0,
S3 v1|1 + r2 = 0,

9.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

379

for the longitudinal vibrations, where


S1 = s1 + 2,

S2 = s2 ,

S3 = s3 + 2,

and the equations


u| = 12 (1|1 + u|11 ),
1
1| = 12 (1 + u|1 ) + (N + 2)1|11 ,
12
and boundary conditions
1 + u|1 = 0,
1|1 = 0,
for the flexural vibrations.
Calculation of the spectra. The frequency equations of these uncoupled
systems can be derived by the standard procedure. For the symmetric longitudinal vibrations the frequencies are sought as the roots of the determinantal
equation
det Cij = 0,

(9.71)

where i, j run from 1 to 3 and


C1j = cos j l,
C2j = j (S2 aj + r3 bj ) sin j l,
C3j = (S3 2j bj + r2 aj ) cos j l.
Here 2i are the three roots of the cubic equation
(S1 2 2 )(S2 2 2 + 22 )(S3 2 2 + 32 )
2 2
(S3 2 2 + 32 )r12 2 (S1 2 2 )r23
= 0,

(9.72)

while ai , bi are given by


S1 2i 2
,
r1
r23 ai
.
bi =
2
S3 i + 32 2

ai =

For the symmetric flexural vibration the frequency equation reads


det Bij = 0,

(9.73)

380

CHAPTER 9. PIEZOELECTRIC SHELLS

where i, j now run from 1 to 2 only, and


1
sin j d,
j
B2j = (2 12 2j ) cos j d.
B1j =

Here 2i are the roots of the quadratic equation


(12 2 2 )(

N + 2 2
+ 12 2 ) 2 14 2 = 0.
12

(9.74)

Figure 9.4: Frequency spectra of barium titanate strips with free edges.
The coefficients of the 2-D equations are taken from the table at the end of
Section 9.3. The determinants in (9.71) and (9.73) are evaluated by exactly
the same numerical procedures as applied to the frequency equations obtained
from (7.113). The computation depends on the behaviour of i , the roots of
the dispersion equations (9.72) and (9.74). In Figure 9.4 the eigenfrequencies
of these systems are plotted against the dimensionless length l = L/h of
the shell. The intersection points of the two sets of curves correspond to
the multiple frequencies in the degenerate case h? = 0. Now let h? be a
small number differing from zero. Taking the terms of oder h? in (9.68)
and (9.69) into account, we obtain the coupled system with small coupling
between u1 , , v1 and u, 1 . According to the theory of perturbations [45] the
multiplicity of the frequencies should disappear. Obviously, the solid lines
in Figure 9.4 change to the dashed lines, which, for the typical intersection
point, look like that shown in Figure 7.21. Thus, the high-frequency spectra

9.5. FREQUENCY SPECTRA OF CYLINDRICAL SHELLS

381

of piezoceramic cylindrical shells (in the axisymmetric case) differ from those
of piezoceramic strips only near intersection points by the mode switching
effect.5 Particularly, the edge resonance vibrations are also observed in semiinfinite piezoceramic cylindrical shells. For barium titanate cylindrical shells
the edge resonance frequency is equal to e = 4.37, which is the same as that
of the semi-infinite barium titanate strip.

One sees that this method does not work in the low-frequency range, since the error
is comparable with the eigenfrequencies.

382

CHAPTER 9. PIEZOELECTRIC SHELLS


Problems

1. Derive the equations determining high frequencies of closed cylindrical


shells with the free edges for n = 1.
2. Determine the high-frequency spectra of closed cylindrical shells made
of PZT-5.
3. Determine the high-frequency spectra of open piezoceramic cylindrical
shells.

Chapter 10
Piezoelectric rods
10.1

One-dimensional equations

Variational principle. Piezoelectric rods, widely used in physical acoustics


as resonators, often work in the high-frequency regime so that the equations
of vibrations derived in Chapter 6 become inapplicable in such situations.
This Chapter is devoted to the construction of a one-dimensional theory of
piezoceramic rods with longitudinal polarization that enables one to determine their resonant vibrations in a wide range of frequencies. For simplicity
we assume that the rod, in its natural state, is straight and untwisted and its
cross section is bounded by a circle of diameter h. The lateral boundary of
the rod is not electroded, while its ends are covered with a pair of electrodes,
and the values of the electric potential are specified on them. Taking the
variational approach as the basis, we postulate that the resonant vibrations
of the rod occur in accordance with the following variational principle
Z

t1

( ) dx dt = 0,

J =
t0

(10.1)

where x is the co-ordinate along the rod axis, and are the 1-D kinetic
energy and electric enthalpy density, respectively. The functional (10.1) is
called, as before, the 1-D action functional.
As the frequency increases, it is natural to assume that some internal
degrees of freedom corresponding to branches of the rod thickness vibrations
will be excited and become more and more involved, so that they should be
included as unknown functions in the 1-D kinetic energy and electric enthalpy
densities. Within this approach we have to specify i) the list of mostly
essential external and internal degrees of freedom, and ii) the 1-D kinetic
energy and electric enthalpy densities depending on these degrees of freedom.
383

384

CHAPTER 10. PIEZOELECTRIC RODS

Since we are concerned here with the resonant vibrations, whose frequencies
are among mechanical eigenfrequencies under short-circuit conditions, the
electric potential at both edges of the rod can be set equal to zero. The
essence of the proposed theory of high-frequency longitudinal vibrations of
piezoceramic rods is expressed by the following formulae
1
= h2 (u 2 + 2 + v 2 ),
2

(10.2)

1
= Gh2 [s1 (u0 )2 + s2 ( 0 )2 + s3 (v 0 )2 + 2r1 h1 u0
2
+2r2 h1 v 0 + 2r3 h1 v 0 + 22 h2 2 + 32 h2 v 2 ].

(10.3)

and

In (10.2) and (10.3) u(x, t) is the average longitudinal displacement (with


some weight), (x, t) and v(x, t) describe the two thickness-stretch and thickness-shear branches of longitudinal vibrations, and the prime is used to denote the derivative with respect to x. It is assumed that the three unknown
functions u, , v are continuous and have continuous derivatives with respect
to x and t. The density of the 1-D kinetic energy represent a simplest
quadratic form of u,
and v that is positive definite. In the density of the
1-D electric enthalpy there are no terms of the type u2 , 0 u, u0 v, uv 0 , u0 v 0 ,
u0 0 . Since the electric potential depends on v linearly, the variation of v
should vanish at the boundary x = 0, L.
The construction of the theory is completed by specifying the constants
G, s1 , s2 , s3 , r1 , r2 , r3 , 2 , 3 and indicating the way to restore the 3-D
displacement and electric potential field of the rod from the functions u,
and v. The next three sections will be devoted to this problem.
Boundary-value problem. It turns out that the 1-D kinetic energy and
electric enthalpy densities (10.2) and (10.3), up to the constants, are identical
in form with the corresponding densities (8.2) and (8.3) for elastic rods.
Therefore the equations of motion for piezoceramic rods can immediately be
derived

u = G(s1 u00 + r1 h1 0 ),
= G(s2 00 r1 h1 u0 r23 h1 v 0 2 h2 ),
2

00

v = G(s3 v + r23 h

(10.4)

32 h2 v),

where r23 = r2 r3 . Taking into account the above-mentioned constraints


on v, we obtain for the traction-free edges of the rod the following boundary

10.2. LONG-WAVE ASYMPTOTIC ANALYSIS

385

conditions at x = 0, L
s1 u0 + r1 h1 = 0,
0 = 0, v = 0.

(10.5)

The boundary conditions of this type differ from those for elastic rods. We
will see later how these boundary conditions influence the frequency spectra of free rods. The formulation of the boundary-value problem should be
completed by specifying the initial conditions at t = t0
u|t0 = u0 ,
u|
t0 = u1 ,

|t0 = 0 ,
t = 1,
|
0

v|t0 = v 0 ,
v|
t0 = v 1 .

Problems
1. Establish the dispersion relation for waves propagating in the longitudinally polarized piezoceramic rod.
2. Prove that if s1 > 0, s2 > 0, s3 > 0, the system (10.4) is hyperbolic.
3. Formulate the variational principle for the eigenvalue problem and
prove the orthogonality of the eigenfunctions with respect to the energy.

10.2

Long-wave asymptotic analysis

3-D variational formulation. We first formulate the problem of resonant


vibrations of the piezoceramic rod that is straight and untwisted in its natural
state. Consider the domain B E specified by
B = {(x1 , x2 , x3 )|(x1 , x2 ) S, x3 x (0, L)},
where {xi } is a cartesian co-ordinate system (the subscript 3 is usually omitted for short). For simplicity we assume that S is bounded by a circle of
diameter h and that h  L. A linear piezoelectric body occupying the domain B in its stress-free undeformed state is called a piezoelectric rod. We
assume that the material of the rod is transversely isotropic about its central
axis (piezoceramic rod with longitudinal polarization). The lateral surface of
the rod is not electroded, while its ends are covered by a pair of electrodes,
and the values 0 /2 are specified on them. Since we are concerned with

386

CHAPTER 10. PIEZOELECTRIC RODS

the resonant vibrations, 0 can be set equal to zero. According to the variational principle (2.45), (2.47) the true displacement field w
i and the electric
potential satisfy the variational equation
Z t1 Z L Z
(T W ) da dx dt = 0,
(10.6)
I =
t0

under the short-circuit conditions


|x=0,L = 0.

(10.7)

Here T is the kinetic energy density


1
1
T = w i w i = (w 2 + w 2 ),
2
2
and W is the electric enthalpy density, which, for the piezoceramic rod with
longitudinal polarization, is given by
G
W = [(w, )2 + 2w(,) w(,) + (w, + w,x )2
2
+ c(w,x )2 + 2w, w,x + 2dw, ,x + 2ew,x ,x
+ 2f (w, + w,x ), (, )2 (,x )2 ].
Since the co-ordinate system is cartesian, we can lower all indices, keeping in
mind the summation convention. The electroelastic constants are given by
the formulae (9.7). The variations of w , w should vanish at t = t0 , t1 and at
the rod ends x = 0, L which are assumed to be clamped.
By stretching the transverse co-ordinates with = x /h we make the
electric enthalpy depend on h explicitly
W =

G
[h2 (w| )2 + 2h2 w(|) w(|) + (h1 w| + w,x )2
2
+ c(w,x )2 + 2h1 w| w,x + 2dh1 w| ,x + 2ew,x ,x
+ 2f (h1 w| + w,x )h1 | h2 (| )2 (,x )2 ]. (10.8)

Here and in what follows the vertical bar preceding a Greek index denotes
the partial differentiation with respect to , while h.i denotes the integral
over S = { |x S}.
The first step of the variational-asymptotic procedure. We analyze
the variational equation (10.6) in the long-wave range, for which the parameter h/l is small everywhere inside the rod, with l the characteristic wavelength

10.2. LONG-WAVE ASYMPTOTIC ANALYSIS

387

in the longitudinal directions. Discarding formally all small terms in (10.6)


we obtain at the zero approximation the equation
2

t1

h
t0

1
G
h (w 2 + w 2 ) [h2 (w| )2 + 2h2 w(|) w(|)
2
2
2
2
+ h (w| ) + 2f h2 w| | h2 (| )2 ]i dx dt = 0. (10.9)

Restricting ourselves to the piezoactive longitudinal vibrations, we can see


that the solutions of the equation (10.9) fall into two series of vibrations.
The series of thickness-shear vibrations, denoted Lk , are characterized by
w = vq( ),

= v ( ),

w = 0,

(10.10)

where q( ), ( ) correspond to the even solutions of the variational problem


hq| q| + f q| | + f | q| | | k2 qqi = 0,

(10.11)

The series of thickness-stretch vibrations, denoted L , are characterized by


w = = 0,

w = p ( ),

(10.12)

where p ( ) are odd solutions of the variational problem


hp| p| + 2p(|) p| 2 p p i = 0.

(10.13)

The eigenvalues and k run through a countable set of values; however,


indices are attached neither to them nor to the corresponding eigenfunctions
in order to avoid complicated notations. Since the solutions of the variational
problems (10.11) and (10.13) are determined uniquely up to a constant factor,
the following normalization conditions can be imposed
hq 2 i = hp2 i = 1.

(10.14)

They are chosen so as to simplify later the average kinetic energy. Further,
the functions v, depend harmonically on t with frequency which is determined from the appropriate values of k or according to
k =

k cs
h

or =

cs
,
h

(10.15)

p
where cs = G/. For functions v, independent of x, each of the solutions
given above represents an exact solution of the equations of 3-D piezoelectricity for an infinite rod and corresponds to the synchronized vibrations of

388

CHAPTER 10. PIEZOELECTRIC RODS

transverse cross sections along the rod (with the zero longitudinal wave number). The frequencies (10.15) will be called cut-off frequencies. For vibrations
whose amplitude and frequency vary slowly in the longitudinal directions of
the rods, the equations (10.10) and (10.12) can be regarded as the zero approximations, where
2
v k2 v,
,
(10.16)
except that corresponding to the zero cut-off frequency. The displacement
and electric potential of this classical low-frequency branch of longitudinal
vibrations, in the first step of the variational-asymptotic procedure, turn out
to be independent from (see Chapter 6).
The second step of the variational-asymptotic procedure. At this
step we find the next refinement for the displacements and electric potential
of each individual branch of vibrations near the cut-off frequencies. Consider
a branch belonging to the series Lk . Regarding v as a given function of x
and t satisfying the constraints
v = 0 at x = 0, L,

(10.17)

we seek w for that branch. Keeping the principal terms depending on w


and the principal cross terms between w and v in (10.6), we obtain the
variational equation
Z

t1

1
1
h k2 w2 G[h2 (w| )2 + 2h2 w(|) w(|)
2
2
t0
0
1 0
1 0
2h v q| w + 2h v qw| + 2dh1 v 0 w| 2f h1 v 0 | w ]i dx dt = 0.
2

The first, fourth and last terms of this equation are obtained after the integration by parts with respect to t and x, respectively. In the long-wave range
near the cut-off frequency the term w can then be replaced by k2 w . The
terms that go to the boundaries x = 0, L vanish due to (10.17) which is also
consistent with the conditions (10.7). From this equation it follows that the
displacements and the electric potential of the branch Lk are given by
w = vq( ),

w = hv 0 s ( ),

= v ( ),

(10.18)

where s ( ) is the solution of the following variational problem


h(s| )2 + 2s(|) s(|) 2q| s + 2qs|
+2d s| 2f | s k2 s2 i = 0.

(10.19)

10.2. LONG-WAVE ASYMPTOTIC ANALYSIS

389

We now turn to a branch belonging to the series L . Regarding as a


given function of x and t satisfying the constraints
= 0 at x = 0, L,

(10.20)

we seek w and for that branch. We impose on the following constraint


hi = 0,
that is, the average electric potential vanishes. According to the variationalasymptotic procedure we must retain the principal terms containing w,
and the principal cross terms between w, and in (10.6). As a result we
obtain the following variational equation
Z

t1

1 2 2 1
w G[h2 (w| )2 + 2h1 0 p w| 2h1 0 p| w
h
2
2
t0
0
2dh1 0 p| + 2f h2 w| | + 2f h1 0 p | h2 (| )2 ]i dx dt = 0.
2

The first and fourth terms in this equation were obtained by the integration
over t and x by parts taking the boundary condition (10.20) into account.
Hence the following asymptotic formulae hold true for the distributions of
displacements and electric potential of the branch L
w = p ( ),

w = h 0 r( ),

= h 0 ( ),

(10.21)

where r( ), ( ) is the solution of the following variational problem


h(r| )2 + 2p r| 2p| r 2dp| + 2f r| |
+2f p | (| )2 2 r2 i = 0,

(10.22)

which must be subject to the constraint


hi = 0.

(10.23)

Average Lagrangians of individual branches of vibrations. We represent the displacements w, w and the electric potential in form of the
infinite series of branches Lk and L given above, where v, are now arbitrary functions of x and t satisfying the constraints (10.17) and (10.20)
at the boundaries of the rod. After substituting these series into the action
functional and integrating over the cross section we neglect those small terms
of order h/l compared with 1. It turns out that the thickness branches are

390

CHAPTER 10. PIEZOELECTRIC RODS

orthogonal relative to the action functional in the long wave range, provided
the shell edge is clamped. Therefore the average functional has the form
Z t1 Z L
2
dx dt,

(10.24)
J =h
t0

is the sum of average Lagrangians of low


where the average Lagrangian
frequency and thickness branches.
Consider first the classical branch of low-frequency longitudinal vibrations, whose displacements and electric field are given by
w = ua,

w = hu0 m ( ),

= 0,

(10.25)

where u describe the


mean longitudinal displacement of the rod, the constant
1/2 = 2/ being chosen to simplify the kinetic energy, and m
a = |S|
satisfy the variational equation
h(m| )2 + 2m(|) m(|) + 2am| i = 0.

(10.26)

Substituting (10.25) into the 3-D action functional, retaining the principal
terms and averaging over the cross section, we obtain
G
= 1 (u)
2 c(u0 )2 ,

2
2
where the constant c is equal to
c = c + ham| i.
Consider a branch in the series of thickness-shear vibrations Lk , whose
displacements and electric potential are given by the asymptotic formulae
(10.18). Substituting (10.18) into the 3-D action functional, keeping the
principal terms and averaging over the cross section, we obtain the following
average Lagrangian
= [v 2 + l2 h2 (v 0 )2 ] G [k2 h2 v 2 + l1 (v 0 )2 ],

2
2

(10.27)

where the coefficients l1 , l2 are given by


l1 = h(s| )2 + 2s(|) s(|) 2q| s + cq 2 + 2s| q
+ 2ds| + 2eq 2f s | 2 i, l2 = hs2 i.
Since v describes harmonic vibrations in the long-wave range near the cut-off
frequency k , for which the estimation (10.16) holds in the first approximation, the small correction term (1/2)l2 h2 (v 0 )2 in (10.27) can be replaced

10.2. LONG-WAVE ASYMPTOTIC ANALYSIS

391

by (1/2)l2 k2 h2 (v 0 )2 , which is equal to (1/2)Gl2 k2 (v 0 )2 . Then the principal


terms of the branch Lk in the average Lagrangian are given by the following
final formula
= 1 v 2 1 [2 h2 v 2 + kk (v 0 )2 ].
(10.28)

2
2
where, with the account of the variational equation (10.19), we obtain for kk
kk = hcq 2 + 2eq 2 i + hq| s + s| q
+ds| + 2eq f s | i.

(10.29)

Finally, we examine a branch in the series of thickness-stretch vibrations


L , whose displacements and electric potential are given by the asymptotic
formulae (10.21). Substituting (10.21) into the 3-D action functional, neglecting small terms and averaging over the cross section, we obtain the
following average Lagrangian
= [ 2 + l2 h2 ( 0 )2 ] G [h2 2 2 + l1 ( 0 )2 ],

2
2
where the coefficients are of the form
l1 = h(r| + p )2 2p| r 2dp|
+ 2f (r| + p )| 2| i, l2 = hr2 i.
Within the first-order approximation we can further simplify the average
Lagrangian. Indeed, by integrating the terms l2 h2 ( 0 )2 over t by parts and
2
, the average Lagrangian for this branch becomes
replacing by
= 2 G [h2 2 2 + k ( 0 )2 ].

2
2

(10.30)

Making use of the variational equations (10.22), we can simplify the expressions for k yielding
k = 1 + hp r| p| r dp| + f p | i.

(10.31)

By varying the 1-D action functional (10.24) with the average Lagrangians
from (10.28) and (10.30) one can easily derive the equations of high-frequency
long-wave vibrations, which are similar to those for elastic rods. Note that for
some range of material parameters the coefficient k may become negative or
infinite, which demonstrates still the true behaviour of the dispersion curves
in the long-wave range, but leads to the ill-posedness of the corresponding
boundary-value problem. This feature could be removed by extrapolating the
theory to short waves taking into account the cross terms between branches,
as we shall see in the next section.

392

CHAPTER 10. PIEZOELECTRIC RODS


Problems

1. Prove that the solution of (10.26) is given by


m =

a
2
c = c
.
2( + )
+

2. Derive the equations and boundary conditions for q and from (10.11)
and for p from (10.13).
3. Do the same as in the previous problem for s and r, .

4. Prove the orthogonality between different thickness branches with respect to the energy in the long-wave range, provided the edges of the
rod are clamped.

10.3

Short-wave extrapolation

Displacements and electric field. Let us consider free vibrations of the


piezoceramic rod under short-circuit conditions, with traction free boundary
conditions at its edges. The vibrations will be regarded as the superposition
of the branches Lk (0), L (0), Lk (1). The branch Lk (0), in the long-wave
range, describes the low-frequency vibrations, the other ones the thickness
vibrations with the lowest cut-off frequencies. The dynamic equations contain
three unknown functions of the longitudinal co-ordinate, x, and the time
t), v(x, t) (the symbols without the bar are reserved for the
t: u(x, t), (x,
functions in the final equations).
Thus, we represent the displacements and the electric potential of the
piezoceramic rod in the form
w = ua + vq + h10 r,
+ h
w = p
u0 m + h
v 0 s ,
= v + h0 .

(10.32)

In this equation the function u corresponds to the low-frequency longitudinal


branch of the vibrations and describes the mean longitudinal displacement
of the rod (with some weight). The functions v and correspond to the first
longitudinal thickness branches in the series Lk and L . The basis functions
q, and p are even and odd eigenfunctions in variational problems (10.11)
and (10.13), and the functions s and r, are determined in terms of q, and
p as the solutions of problems (10.19) and (10.22), respectively. Since the

10.3. SHORT-WAVE EXTRAPOLATION

393

electric potential that depends linearly on v should satisfy the short-circuit


conditions, we require that v satisfy the following kinematical constraints
v = 0 at x = 0, L.
We substitute (10.32) into the 3-D action functional (10.6) and integrate
Keeping the principal terms of each branch and the
over the cross section S.
principal cross terms in the average Lagrangian and taking into account the
results in the previous section, we obtain
= 1 (u 2 +2b1 hu 0 + 2 +2b2 h u 0 +2b3 h v 0 + v 2 +2b4 hv 0 ) 1 G[ 2 h2 2

2
2
2
v 0 + 2r3 h1 v0
v 0 + 2r1 h1 u0 + 2r2 h1
u0 + 2a2 h1
+ 2a1 h1
+ 32 h2 v2 + 2a3 h1 v0 + c(
u0 )2 + k2 (0 )2 + k3 (
v 0 )2 ]. (10.33)
Here 2 and 3 are the first eigenvalues and k of the longitudinal thickness
vibrations in problems (10.11) and (10.13). The formulae for the remaining
coefficients have the form
b1 = hari, b2 = hp m i, b3 = hp s i, b4 = hqri,
a1 = hp| m| + 2p(|) m(|) i,
a2 = hp| s| + 2p(|) s(|) i,
a3 = hq| r| + f q| | + f r| | | | i,
r1 = hp| ai, r2 = hp| q + dp| i,
r3 = hq| p + f p | i.

(10.34)

The coefficients k2 , k3 are given by the formulae (10.31) and (10.29), respectively.
It turns out that the following relationships between the coefficients hold
true
b1 = b2 , b3 = b4 , a1 = b2 22 , a2 = b3 22 , a3 = b4 32 ,
r1 = b1 22 = a1 , r2 r3 = b3 (32 22 ) = a3 a2 .

(10.35)

This can be proved with the help of the variational problems (10.11), (10.13),
(10.19), (10.22) and (10.26), without calculating the coefficients. Indeed, the
substitutions p = m for p in (10.13) and r = a, = 0 for r, in
(10.22) lead to
hp| m| + 2p(|) m(|) i = 22 hp m i = 22 b2 ,
hp| ai = 22 hrai = 22 b1 .

(10.36)
(10.37)

394

CHAPTER 10. PIEZOELECTRIC RODS

From (10.34), (10.36) and (10.37) the equalities a1 = 22 b2 and r1 = 22 b1


follow. Substracting (10.37) from (10.36), we obtain the equality
hp| m| + 2p(|) m(|) + ap| i = 22 (b2 b1 ).
The left-hand side of this coincides with the variational equation (10.26) for
m , when we substitute therein m = p . Therefore 22 (b2 b1 ) = 0, or
b2 = b1 .
The substitutions q = r, = in (10.11), p = s in (10.13), s = p
in (10.19) and r = q, = in (10.22) yield
hq| r| + f q| | + f | r|
| | i = 32 hqri = 32 b4 ,
hp| s| + 2p(|) s(|) i = 22 hp s i = 22 b3 ,
hs| p| + 2s(|) p(|) + qp| q| p
+d p| f | p i = 32 hs p i = 32 b3 ,

(10.38)
(10.39)
(10.40)

and
hr| q| + p q| p| q dp| + f r| |
+f | q| + f p | | i = 22 hrqi = 22 b4 ,

(10.41)

respectively. From (10.38) and (10.39) follow a3 = 32 b4 and a2 = 22 b3 .


Substracting (10.39) from (10.40), we get the equality
hq| p + qp| + d p| f | p i = b3 (32 22 ).
According to (10.34) this equality means r2 r3 = b3 (32 22 ). Finally, adding
(10.40) and (10.41) together and substracting (10.38) and (10.39) from the
result we get
(22 32 )(b4 b3 ) = 0.

(10.42)

Since 2 6= 3 , it follows from (10.42) that b3 = b4 . Thus, relations (10.35)


are proved.
It follows from relations (10.35) that the principal cross terms in the average Lagrangian (10.33) form divergent terms (null Lagrangian) that do not
affect the equations of vibrations in the long-wave range. Consequently, a
method of short-wave extrapolation is possible in which all the cross terms in
(10.33) are discarded, which would result in uncoupling between the vibration branches. However, additional analysis shows that such extrapolation
leads to a qualitatively false description of the dispersion curves and integral

10.3. SHORT-WAVE EXTRAPOLATION

395

characteristics of the rod in the short-wave range (cf. Section 8.5). Therefore,
we will here keep all the cross terms and just try to seek the change of the
unknown functions that would simplify (10.33), i.e., would reduce (10.33) to
an expression without second and higher-order derivatives with respect to x
and t.
Using (10.35) we transform the Lagrangian (10.33) to
= 1 [(u + b2 h 0 )2 + ( + b2 hu 0 + b3 hv 0 )2 + (v + b3 h 0 )2

2
1
(b22 + b23 )h2 ( 0 )2 b22 h2 (u 0 )2 b23 h2 (v 0 )2 ] G[22 h2 ( + b2 h
u0 + b3 h
v 0 )2
2
v 0 + 2r3 h1 v0 + 2 h2 (
+ 2r1 h1 u0 + 2r2 h1
v + b3 h0 )2
3
+ (
c b22 22 )(
u0 )2 + (k2 b23 32 )(0 )2 + (k3 b23 22 )(
v 0 )2 ]. (10.43)
When transforming (10.33) to (10.43) the cross terms h2 u 0 v 0 and u0 v0 are
neglected as small compared with other cross terms. In (10.43) the terms in
the kinetic energy 1/2(b22 + b23 )h2 ( 0 )2 and 1/2b23 h2 (v 0 )2 at long waves
v 0 )2 . The term
can be replaced by 1/2(b22 + b23 )22 (0 )2 and 1/2b23 (
1/2b22 h2 (u 0 )2 is small at long waves and can be omitted. The Lagrangian
then has the form
= 1 [(u + b2 h 0 )2 + ( + b2 hu 0 + b3 hv 0 )2 + (v + b3 h 0 )2 ]

2
1
v 0
G[22 h2 ( + b2 h
u0 + b3 h
v 0 )2 + 2r1 h1 u0 + 2r2 h1
2
+ 2r3 h1 v0 + 32 h2 (
v + b3 h0 )2 + (
c b22 22 )(
u0 )2
+ (k2 b23 32 + (b22 + b23 )22 )(0 )2 + (k3 b23 22 + b23 32 )(
v 0 )2 ]. (10.44)
In order to search for a short-wave extrapolation which does not contain
second and higher derivatives in the Lagrangian let us make the changes of
unknown functions u u, and v v, where
u = u + b2 h0 ,
= + b2 h
u0 + b3 h
v0,
v = v + b3 h0 .
We require that the new unknown function v satisfy the constraints
v = 0 at x = 0, L.

396

CHAPTER 10. PIEZOELECTRIC RODS

Keeping the principal terms of u, , v in (10.44), we arrive at


= 1 (u 2 + 2 + v 2 ) 1 [22 h2 2 + 32 h2 v 2 + 2r1 h1 u0

2
2
2r1 b2 (u0 )2 2r1 b2 00 + 2r2 h1 v 0 2r2 b3 00 2r2 b3 (v 0 )2
u0 )2
+ 2r3 h1 v 0 2r3 b3 ( 0 )2 2r3 b3 vv 00 + (
c b22 22 )(
+ (k2 b23 32 + (b22 + b23 )22 )(0 )2 + (k3 b23 22 + b23 32 )(
v 0 )2 ].
By adding a null Lagrangian, we can replace terms of the type 00 and
vv 00 by ( 0 )2 and (v 0 )2 , respectively, and derive finally the formulae (10.2)
and (10.3) for and , where the coefficients s1 , s2 , s3 are given by
s1 = c b22 22 2r1 b2 ,
s2 = k2 b23 32 + (b22 + b23 )22 + 2r1 b2 + 2r2 b3 2r3 b3 ,
s3 = k3 b23 22 + b23 32 2r2 b3 + 2r3 b3 .
Making use of relations (10.35) again, we obtain
r12
,
22
r2
r2
s2 = k2 12 + 2 23 2 ,
2
3 2
2
r
s3 = k3 2 23 2 .
3 2
s1 = c +

(10.45)

Thus, the formulae of the 1-D theory of high-frequency longitudinal vibrations is justified by the short-wave extrapolation procedure, which consists of
the operations of adding or removing a) terms that are small in the long-wave
range, or b) null Lagrangians. Besides, by solving the cross section problems
we can calculate all the coefficients of the theory.
Problems
1. Prove that the short-wave extrapolations proposed are asymptotically
equivalent to the theory of vibrations of piezoceramic rods in the lowfrequency limit.

2. Derive the 1-D equations describing the antiresonant vibrations of the


piezoceramic rod with the same electrode arrangement.

3. Derive the 1-D equations of high-frequency resonant vibrations for sideelectroded rods polarized along the thickness direction.

10.4. CROSS SECTION PROBLEMS

10.4

397

Cross section problems

Determination of the eigenfunctions. We look for the axially symmetric solutions of the cross section problems. Let us introduce the polar
co-ordinates %, . For the axially symmetric vibrations obviously p s 0,
while p% , r, and q, , s% depend only on %. Therefore, variational problems
(10.11), (10.13), (10.19), (10.22) reduce to boundary-value problems for ordinary Bessel differential equations and are solved explicitly in terms of Bessels
functions.
We first determine the functions p of the branch L (0) satisfying variational problem (10.13). Varying (10.13), we obtain the Euler equation
( + )p| + 2 p = 22 p ,

(10.46)

and the natural boundary condition

p| + 2p(|) = 0 at S.

(10.47)

Seeking p in the form p = p| , where p depends only on %, we reduce


(10.46) to the following equation in the polar co-ordinates
1 d dp
(% ) + 22 p = 0,
% d% d%

22 =

22
.
+ 2

(10.48)

The non-singular solution of (10.48) is given by


p = P J0 (2 %),

p% = P 2 J1 (2 %).

(10.49)

From (10.47) we obtain the following boundary condition for p

1 d dp
d2 p
(% ) + 2 2 = 0,
% d% d%
d%

at % = 1/2.

(10.50)

Substitution of (10.49) into (10.50) leads to the frequency equation for 2


4e21 J1 (2 /2) = 2 J0 (2 /2),

e21 =

.
+ 2

The constant P is determined from the normalization condition


Z 1/2
2
p2% %d% = 1.
hp i = 2
0

With p% from (10.49) we find that


P =

21
,
2 J0 (2 /2)

2e1
1 = p
.
2
(2 /e1 ) /4 4(1 e21 )

398

CHAPTER 10. PIEZOELECTRIC RODS

Having found p , we can determine r, by solving the variational problem


(10.22) under the constraint (10.23). Introducing the Lagrange multiplier
and varying (10.22), we obtain the system of equations
2 r + (1 + )p| + f 2 + 22 r = 0,
f 2 r + (d + f )p| 2 = ,

(10.51)

and the boundary conditions at S


(r| + p + f | ) = 0,
(f r| + f p | ) = 0.

(10.52)

Expressing 2 in the second equation of (10.51) through the other quantities, substituting the result into the first equation, and recalling that p| =
2 p, we obtain for r
2 r + 2 p + 22 r =

f
,
+ f2

(10.53)

where
1 + + f (d + f )/
,
=
1 + f 2 /

22

22
.
=
1 + f 2 /

Analogously, from (10.52) we derive the following boundary condition for r


at S
(r| + p| ) = 0.

(10.54)

Rewriting (10.53) and (10.54) in the polar co-ordinates, we obtain


1 d dr
f
(% ) + 22 r = P 22 J0 (2 %)
,
% d% d%
+ f2
dr
= P 2 J1 (2 /2) at % = 1/2.
d%

(10.55)

Problem (10.55) yields the following solution


r = R1 + R2 J0 (2 %) + R3 J0 (2 %),
where the constants R1 , R2 , R3 are equal to
f
P 22
,
R
=
,
2
( + f 2 )22
22 22
(R2 + P )2 J1 (2 /2)
R3 =
.
2 J1 (/2)

R1 =

(10.56)

10.4. CROSS SECTION PROBLEMS

399

Now we substitute r into the second equations of (10.51) and (10.52) to


obtain the boundary-value problem for
f 2
d+f 2

r+
p+ ,

| = 0.
2 =

(10.57)
(10.58)

The solution of (10.58) gives


=

f
d+f
%2
r+
p+
+ a,

where
= 4dP 2 J1 (2 /2),
and a is a constant which should be determined from the condition
hi = 0.
Substituting for p and r their expressions (10.49) and (10.56), we represent
in the form
= 1 + 2 J0 (2 %) + 3 J0 (2 %) + 4 %2 ,
where
2 =

f
d+f
R2 +
P,

3 =

f
R3 ,

4 =

,
4

and
1 = 2

4J1 (2 /2)
4J1 (2 /2)
3
4 /8.
2
2

We now turn to the functions q, , which should satisfy variational problem (10.11). The Euler equations of (10.11) read
2 q + f 2 32 q = 0,
f 2 q 2 = 0,
while the natural boundary conditions at S are of the form
(q| + f | ) = 0,
(f q| | ) = 0.

400

CHAPTER 10. PIEZOELECTRIC RODS

It is easy to see that


=

f
q.

Since q depends only on %, the boundary-value problem for it in the polar


co-ordinates reduces to
1 d dq
32
(% ) + 32 q = 0, 32 =
,
% d% d%
1 + f 2 /
dq
= 0 at % = 1/2,
d%

(10.59)

It follows from (10.59) that


q = QJ0 (3 %),

(10.60)

where 3 is the lowest positive root of


J1 (3 /2) = 0 3 = 7.6634.
Using the normalization condition (10.14) it is easy to show that

Q = 2/ J0 (3 /2).
Knowing the functions q, , we can substitute them into the variational
problem (10.19) to derive the equations for s . Seeking s in the form s =
s| , where s depends only on %, the following boundary-value problem for s
is obtained
1 + + (d + f )f /
1 d ds
(% ) + 32 s =
q,
% d% d%
+ 2
1 d ds
d2 s

(% ) + 2 2 = ( + df /)q at % = 1/2,
% d% d%
d%

(10.61)

where 32 = 32 /( + 2). With q from (10.60), problem (10.61) yields the


following solution for s
s = S1 J0 (3 %) + S2 J0 (3 %),
where
1 + + (d + f )f /
Q,
( + 2)(32 32 )
2 (1 + + (d + f )f /)32 /(32 32 ) + ( + df /)
S2 =
.
32 J0 (3 /2) 43 J1 (3 /2)

S1 =

10.4. CROSS SECTION PROBLEMS

401

Consequently, s% is given by
s% = S1 3 J1 (3 %) S2 3 J1 (3 %).
Calculation of the coefficients. Having found all the solutions of the cross
section problems, we now use the formulae (10.29), (10.31) and (10.34) to
calculate the coefficients of the 1-D theory. Note that for arbitrary functions
f (%) and g(%) independent from the angle , the following formula
Z 1/2
hf gi = 2
f (%)g(%)%d%
(10.62)
0

holds true. Substituting the functions a, p% and q, into (10.34) and making
use of the formula (10.62), we obtain the following expressions for r1 , r2 , r3
r1 =

1 2
,
e21

1 23
,
(32 22 )e21
1 2 2
r3 = (1 + f 2 /) 2 3 2 2 .
(3 2 )e1
r2 = ( + f d/)

The coefficients k2 and k3 are determined from (10.31) and (10.29), respectively. After long, but otherwise standard, calculations we arrive at the
following expressions
1
k2 = 1 + 2{P 22 (R2 + f 2 ) [J12 (2 /2) J0 (2 /2)J2 (2 /2)]
8
1
+ P 2 2 (R3 + f 3 )
[2 J0 (2 /2)J1 (2 /2) 2 J0 (2 /2)J1 (2 /2)]
2(22 22 )
J1 (2 /2)
1
+ P 2 (R1 + d1 )
+ P 22 (R2 + d2 ) [J02 (2 /2) + J12 (2 /2)]
2
8
1
[2 J1 (2 /2)J0 (2 /2) + 2 J1 (2 /2)J0 (2 /2)]
+ P 22 (R3 + d3 )
2(22 22 )
J2 (2 /2)
4J2 (2 /2) 2 J3 (2 /2)
+ P d4
P f 4
},
8
2
and
2ef
f2

2 (1 + + (d + f )f /) QS1 32 J02 (3 /2)

+ [32 + ( + df /)32 ]QS2 3 2


J0 (3 /2)J1 (3 /2).
3 32

k3 =c +

402

CHAPTER 10. PIEZOELECTRIC RODS

Finally, the coefficients s1 , s2 , s3 are calculated from c, k2 , k3 and r1 , r2 , r3


according to (10.45).
Problems
1. Show that for the special case of isotropic elastic rods with
= = , = 1, c = + 2, d = e = f = = = 0,
the distributions of displacements over the cross section reduce to those
in Section 8.4.
2. Show that for isotropic elastic rods the coefficients of the 1-D theory
coincide with those in Section 8.4.

10.5

Frequency spectra

Derivation of the frequency equation. Consider the mechanical eigenvibrations of a piezoceramic rod of circular cross section under short-circuit
conditions, which are governed by the equations of motion (10.4). Introducing the dimensionless variables
s
t G
x
, = ,
=
h
h
we rewrite these equations in the form
u| = s1 u00 + r1 0 ,
| = s2 00 r1 u0 r23 v 0 22 ,
v| = s3 v 00 + r23 0 32 v,

(10.63)

where the prime is used to denote the derivative with respect to .


In the case of harmonic vibrations we seek the solution in the form
u = u()ei ,

= ()e
,

v = v()ei ,

where is the dimensionless frequency. Substituting this into (10.63) and


eliminating the common factor ei , we reduce (10.63) to
s1 u00 + r1 0 + 2 u = 0,
s2 00 r1 u0 r23 v0 + (2 22 ) = 0,
s3 v00 + r23 0 + (2 32 )
v = 0.

(10.64)

10.5. FREQUENCY SPECTRA

403

In the same way, the traction-free boundary conditions (10.5) at = l =


L/2h are reduced to
s1 u0 + r1 1 = 0,
s2 0 = 0, v = 0.

(10.65)

It turns out that the solution of (10.64) can be expressed through three scalar
potentials i , i = 1, 2, 3 as follows
u = 01 + 02 + 03 ,
= a1 1 + a2 2 + a3 3 ,
v = b1 01 + b2 02 + b3 03 ,
provided the potentials i satisfy the following equations
00i + 2i i = 0,

i = 1, 2, 3,

(10.66)

where i are the roots of the dispersion relation


(s1 2 2 )(s2 2 2 + 22 )(s3 2 2 + 32 )
2 2
(s3 2 2 + 32 )r12 2 (s1 2 2 )r23
= 0,
and where ai , bi are expressed through i by
s1 2i 2
,
r1
r23 ai
bi =
.
2
s3 i + 32 2

ai =

We restrict ourselves to the symmetric solution of (10.66) given by


i = ci cos i .
v through i
Expressing u, ,
u =

3
X

ci i sin i ,

i=1

3
X

ai ci cos i ,

i=1

v =

3
X
i=1

bi ci i sin i ,

404

CHAPTER 10. PIEZOELECTRIC RODS

and substituting these formulae into the boundary conditions (10.65), we


obtain the system of homogeneous linear equations for ci
3
X

Cij cj = 0,

i = 1, 2, 3,

j=1

where
C1j = cos j l,
C2j = aj j sin j l,
C3j = bj j sin j l.
Thus, the eigenfrequencies of vibrations should be determined as the roots
of the equation
det Cij = 0.

(10.67)

Calculation of the spectra. We illustrate the calculation of the spectra


with the example of the piezoceramic rods of barium titanate BaTiO3 , whose
electroelastic constants sE , d, T are given in Appendix A3. Calculating the
constants cE , e, S and using the formulae (9.7), we find the numerical values
of , , c, , d, e, f, and , to be equal to
= 1.496, = 0.966, c = 3.318, = 1.503,
d = 0.985, e = 3.959, f = 2.6 (1010 C/N),
= 22.511, = 25.26 (1020 C2 /N2 ).
Making use of the formulae derived in the previous section for the coefficients of the 1-D theory, we find that
2 = 7.889, 3 = 8.739,
r1 = 6.634, r2 = 2.743, r3 = 8.305,
s1 = 3.107, s2 = 0.736, s3 = 2.892.
Having all the coefficients, we can calculate the spectra of the eigenvibrations
of the piezoceramic rods under short-circuit conditions by solving numerically
the frequency equation (10.67) for .
Figure 10.1 shows the first seven roots of this equation as functions of
the dimensionless length of rods l. The frequencies computed by using the

10.5. FREQUENCY SPECTRA

405

Figure 10.1: Frequency spectra of the longitudinal vibrations of piezoceramic


rods: a) 1-D theory of high frequency vibrations: solid line and b) 1-D theory
of low-frequency vibrations: dashed line.
theory of low-frequency longitudinal vibrations are also shown (the dashed
hyperbolas). They are given by

n =

1
(n ).
l
2

(10.68)

One can see that the frequencies n (l) computed in accordance with the
frequency equations (10.67) and (10.68) differ litle for . 3, and the hyperbolas (10.68) approach asymptotically the curves obtained from (10.67)
as 0 (l ). For & 3 fairly strong divergence between the curves
obtained from (10.67) and (10.68) is observed. However, the terrace-like
spectra do not exist at the frequencies below the cut-off frequency 2 , and
no edge mode is observed in this range of frequencies. This behaviour is due
to the boundary conditions (10.65), and is specific for this type of electrode
arrangement. However, for side-electroded piezoceramic rod, polarized in the
thickness direction, the theory and experiment show that the edge mode does
exist.
Problems
1. Calculate the coefficients of the 1-D theory for PZT-4 and PZT-5 piezoceramics.

406

CHAPTER 10. PIEZOELECTRIC RODS

2. Calculate the frequency spectra of the longitudinal vibrations of piezoceramic rods with the clamped edges.

Appendix A
Material constants
A.1

Elastic isotropic materials

Linear elastic isotropic materials possess two independent elastic constants.


Depending on specific problems one can prefer, for instance, the pairs , ,
E, or , . The pairs E, and , are very often used in the shell and rod
theories. For dynamic problems we also use the dimensionless ratios and
e. Relations between the elastic constants are given in the following table:

Constant

E,

E
(1+)(12)

2
1

E
2(1+)

(3+2)
+

2(1+2)
1+

2(+)

1+

+2

2
12

2
1

p
/( + 2)

p
(1 2)/(2 2)

p
(1 )/2

407

408

A.2

APPENDIX A. MATERIAL CONSTANTS

Piezoelectric crystals
Electroelastic moduli for the 32 crystal classes

A.2. PIEZOELECTRIC CRYSTALS


Electroelastic moduli for the 32 crystal classes (continued)

409

410

A.3

APPENDIX A. MATERIAL CONSTANTS

Piezoceramic materials

Piezoceramic materials are transversely isotropic and possess therefore five


elastic, three piezoelectric, and two dielectric constants. Depending on specific problems one can prefer, for instance, the constants cE , e, T or sE , d, S .
Relations between these constants, written in the abbreviated indicial notation, are
E E E E
1
E
c11 c12 c13
s11 sE
12 s13
E
E
E
E
cE
= s E
,
12 c11 c13
12 s11 s13
E
E
E
E
E
cE
c
c
s
s
s
13
13
33
13
13
33
E
1
E
E
e31
s11 s12 s13
d31
E
E
e31 = sE

d31 ,
12 s11 s13
E
E
E
e33
s13 s13 s33
d33
E
E
where sE
12 = s11 s66 /2, and

1
d15
1
, cE
e15 = E ,
66 = E ,
E
s55
s66
s55
T
S
T
= 11 d15 e15 , 33 = 33 2d31 e31 d33 e33
cE
55 =

S11

The following table presents experimental data [9] for some piezoceramics
(0 = 8.854 1012 C2 /Nm2 is the dielectric constant of vacuum):
Quantity

PZT-4

PZT-5

BaTiO3

12 2
m /N
sE
11 , 10

12.3

16.4

9.1

sE
13

-5.31

-7.22

-2.9

sE
33

15.5

18.8

9.5

sE
55

39.0

47.5

22.8

sE
66

32.7

44.3

23.6

d15 , 1012 C/N

496

584

260

d31

-123

-172

-78

d33

-289

-374

-190

T11 /0

1475

1730

1450

635

830

1260

7.5

7.75

5.7

T33 /0
3

, 103 kg/m

Appendix B
List of notations
General
Time t
Dimensionless time
Three-dimensional Euclidean space E
Cartesian co-ordinates z i , i = 1, 2, 3
Curvilinear co-ordinates xa , a = 1, 2, 3
Vectors and tensors u, ua , t, tab , . . .
3-D metrics gab , g ab
Kronecker delta ij
Gradient u, ua;b
Time derivative of u u
3-D domain U, B
Boundary of B B
Volume element dv
2-D surface S
2-D co-ordinates x , = 1, 2
First and second fundamental forms of a surface a , b
Mean and Gaussian curvatures of a surface H, K
Covariant derivative on a surface u;
Area element da
1-D curve c(x)
Curvature and torsion of a curve , $
Length element ds or dx
Functional I[w]
Variation of w w
Stationary point w
Frequency and dimensionless frequency ,
411

412

APPENDIX B. LIST OF NOTATIONS

Dimensionless co-ordinates ,
Derivatives with respect to the dimensionless co-ordinates u|
Wave number and dimensionless wave number k,
3-D elasticity
Displacements wa
Strains ab
Stresses ab
Mass density
Elastic moduli C abcd
Elastic moduli for isotropic bodies , or E,
Elastic energy density W (ab )
Kinetic energy density T (w a )
3-D piezoelectricity
Electric potential
Electric field Ea
Electric induction Da
cab
cab ab
, Sab
, S or cabcd
Electroelastic moduli cabcd
D ,h
E ,e
Electric enthalpy W (ab , Ea )
Internal energy U (ab , Da )
Complementary energy (Gibbs function) G( ab , E a )
Elastic enthalpy F ( ab , Da )
Elastic shells
Middle surface S
Thickness h
Characteristic radius of curvature R
Characteristic scale of change of the deformation pattern l
Displacements of the middle surface u , u
Internal degrees of freedom , , v
Measures of extension and bending A , B or A ,
Membrane stresses, bending moments N , M or n , m
2-D kinetic energy density
2-D elastic energy
Elastic rods
Central line c(x)
Cross section S
Diameter of the cross section h

413
Curvatures and torsion , $
Length L
Displacements of the central line u , u
Internal degrees of freedom , 1 , v
Rotation of the cross section
Measures of elongation, bending and twist , ,
Tension, bending and twisting moments T, M , M
1-D kinetic energy density
1-D elastic energy
Piezoelectric shells and rods
2-D
2-D
2-D
2-D
2-D

and
and
and
and
and

1-D
1-D
1-D
1-D
1-D

electric potential
electric field F , F
electric induction G , G
kinetic energy density
electric enthalpy 0 , 1

414

APPENDIX B. LIST OF NOTATIONS

Bibliography
[1] M. Abramowitz and I. A. Stegun. Handbook of Mathematical Functions.
Dover Publications, New York, 1965.
[2] A. E. Armenakas, D. G. Gazis, and G. Herrman. Free vibrations of
circular cylindrical shells. Pergamon Press, New York, 1969.
[3] R. C. Batra and J. S. Yang. Saint-Venants principle in linear piezoelectricity. J. Elasticity, 38:209218, 1995.
[4] V. L. Berdichevsky. On the proof of the Saint-Venant principle for bodies
of arbitrary shape. Appl. Math. Mech. (PMM), 38:851864, 1974.
[5] V. L. Berdichevsky. Variational-asymptotic method for constructing
shell theory. Appl. Math. Mech. (PMM), 43:664687, 1979.
[6] V. L. Berdichevsky. Variational principles of continuum mechanics.
Nauka, Moscow, 1983.
[7] V. L. Berdichevsky and K. C. Le. High frequency, long wave shell vibration. Appl. Math. Mech. (PMM), 44:520525, 1980.
[8] V. L. Berdichevsky and K. C. Le. High frequency vibration of shells.
Soviet Physics Doklady, 27:988990, 1982.
[9] D. A. Berlincourt, D. R. Curran, and H. Jaffe. Piezoelectric and piezomagnetic materials and their function in transducers. In Physical acoustics, volume 1A. Academic Press, New York, 1964.
[10] V. A. Boriseiko, V. S. Martynenko, and A. F. Ulitko. On the theory of
vibrations of piezoceramic shells. Mathematical Physics, 21:7176, 1977.
[11] W. G. Cady. Piezoelectricity. Dover Publications, New York, 1964.
[12] R. Courant and D. Hilbert. Methods of Mathematical Physics, vols. 1
and 2. Interscience, New York, 1953.
415

416

BIBLIOGRAPHY

[13] I. Ekeland and R. Temam. Analyze convexe et problemes variationnels.


Dunod, Paris, 1974.
[14] A. Erdelyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi. Higher
transcendental functions, volume II. McGraw-Hill, New York, 1953.
[15] J. N. Flavin, R. J. Knops, and L. E. Payne. Decay estimates for the
constrained elastic cylinder of variable cross section. Quart. Appl. Math.,
47:325350, 1989.
[16] D. C. Gazis. Three-dimensional investigation of the propagation of waves
in hollow circular cylinders. I. Analytical foundation. J. Acoust. Soc.
Am., 31:568573, 1959.
[17] A. L. Goldenveizer. Theory of thin elastic shells. Nauka, Moscow, 1976.
[18] V. T. Grinchenko and V. V. Meleshko. Harmonic vibrations and waves
in elastic bodies. Naukova Dumka, Kiev, 1981.
[19] M. E. Gurtin. The linear theory of elasticity, volume VIa/2 of Handbuch
der Physik. Springer, Berlin, 1972.
[20] B. Jaffe, W. R. Cook, and H. Jaffe. Piezoelectric ceramics. Academic
Press, New York, 1971.
[21] J. D. Kaplunov. High-frequency stress-strain states. Izv. Akad. Nauk
S.S.S.R, MTT, 25:147157, 1990.
[22] J. D. Kaplunov, L. Y. Kossovich, and E. V. Nolde. Dynamics of thin
walled elastic bodies. Academic Press, New York, 1998.
[23] W. T. Koiter. A consistent first approximation in the general theory
of thin elastic shells. In Proc. IUTAM Symp. Theory of Thin Elastic
Shells, pages 1233, 1960.
[24] W. T. Koiter. On the mathematical foundation of shell theory. In Proc.
Intern. Congress of Math. Nice, 1970, volume 3, pages 123130, Paris,
1971. Gauthier-Villars.
[25] S. S. Kvashnina. High-frequency long-wave vibrations of elastic rods.
Applied Mathematics and Mechanics, 43:335341, 1979.
[26] V. I. Krylov L. V. Kantorovich. Approximate method of higher analysis.
Wiley, New York, 1964.

BIBLIOGRAPHY

417

[27] K. C. Le. High frequency, long wave vibration of piezoelectric ceramic


plates. Soviet Physics Doklady, 27:422423, 1982.
[28] K. C. Le. Fundamental relations of the theory of anisotropic piezoelectric
shells. Bulletin of Moscow University, Math.-Mech., (5):5760, 1984.
[29] K. C. Le. On the edge resonance of the semi-infinite elastic plate. Bulletin of Moscow University, Math.-Mech., (5):5760, 1984.
[30] K. C. Le. High frequency vibrations of piezoelectric ceramic shells. Soviet
Physics Doklady, 30:899900, 1985.
[31] K. C. Le. High frequency longitudinal vibrations of elastic rods. Appl.
Math. Mech. (PMM), 50:335341, 1986.
[32] K. C. Le. The theory of piezoelectric shells. Appl. Math. Mech. (PMM),
50:98105, 1986.
[33] K. C. Le. High frequency vibrations and wave propagation in elastic shells: variational-asymptotic approach. Int. J. Solids Structures,
34:39233939, 1997.
[34] A. W. Leissa. Vibration of shells. NASA, Washington, 1973.
[35] A. E. H. Love. A Treatise on the Mathematical Theory of Elasticity.
Cambridge University Press, Cambridge, 4th edition, 1927.
[36] W. P. Mason. Piezoelectric crystals and their applications to ultrasonics.
D. Van Nostrand Company, New York, 1950.
[37] G. A. Maugin. Continuum mechanics of electromagnetic solids. NorthHolland, Amsterdam, 1988.
[38] G. A. Maugin and A. C. Eringen. Electrodynamics of continua. I
Foundations and solid media. Springer, New York, 1989.
[39] J. C. Maxwell. Electricity and magnetism. Clarendon Press, Oxford,
1892.
[40] A. J. McConnell. Applications of tensor analysis. Dover publications,
New York, 1957.
[41] R. D. Mindlin. The collected papers. Springer, Berlin, 1989.
[42] V. V. Novozhilov. Thin shell theory. Wolters-Noordhoff Publishing,
Groningen, 2th edition, 1970.

418

BIBLIOGRAPHY

[43] J. F. Nye. Physical Properties of Crystals. Clarendon Press, London


and New York, 1957.
[44] I. G. Petrovsky. Lectures on Partial Differential Equations. Dover Publications, New York, 1991.
[45] T. Poston and I. Stewart. Catastrophe Theory and its Applications.
Pitman, Boston, 1978.
[46] W. Prager and J. L. Synge. Approximation in elasticity based on the
concepts of function space. Quart. Appl. Math., 5:121, 1947.
[47] J. W. Rayleigh. The Theory of Sound. Dover Publications, New York,
second edition, 1945.
[48] E. Reissner. Selected works in applied mechanics and mathematics. Jones
& Bartlett Publishers, Boston, 1996.
[49] K. Rektorys. Variational methods in mathematics, science and engineering. D. Reidel Publishing Company, Dordrecht, 1980.
[50] N. N. Rogacheva. The theory of piezoelectric shells and plates. CRC
Press, London, 1994.
[51] M. Yu. Ryazantseva. Flexural vibrations of symmetrical sandwich plates.
Mechanics of Solids (MTT), 20:153159, 1985.
[52] J. L. Sanders. An improved first order approximation theory of thin
shells. In NASA Report, volume 24, 1959.
[53] E. A. G. Shaw. On the resonant vibrations of thick barium titanate
disks. J. Acoust. Soc. America, 30:979984, 1956.
[54] M. Spivak. Differential Geometry. Publish or Perish, Berkeley, 1975.
[55] S. Sternberg. Lectures on Differential Geometry. Chelsea, New York,
1983.
[56] H. F. Tiersten. Linear piezoelectric plate vibrations. Plenum Press, New
York, 1969.
[57] S. Timoshenko. On the transverse vibrations of bars of uniform cross
section. Phil. Mag., 43:125131, 1922.
[58] S. Timoshenko. Vibration Problems in Engineering. D. Van Nostrand
Company, Princeton, 1955.

BIBLIOGRAPHY

419

[59] R. A. Toupin. Saint-Venants principle. Arch. Rat. Mech. Anal., 18:83


96, 1965.
[60] I. A. Vekovisheva. Variational principles in the theory of electroelasticity.
Appl. Mech., 7:129133, 1971.
[61] G. B. Whitham. Linear and Nonlinear Waves. Wiley Interscience, New
York, 1974.
[62] L. C. Young. Lectures on the calculus of variations and optimal control
theory. Chelsea, New York, 1969.

420

BIBLIOGRAPHY

Index
2-D moduli, 172, 216
3-D equations
of motion, 31, 75, 80, 101

Boundary-value problems, 31, 39, 63,


127, 253, 312, 348, 380

Abbreviated indicial notation, 38, 182,


187, 189, 350
Acceleration, 30, 37, 60, 124, 209
Action functional
1-D, 125, 209, 309, 379
2-D, 60, 163, 249, 347
3-D, 32, 40, 68, 130, 214, 257,
314, 350
Antiresonance, 43
Area element, 23, 26
Asymptotic equivalency, 65
Balance equations, 30, 37
Basis, 19, 58, 122, 123, 130
Bending
measures of, 59, 95, 116, 124, 208,
348
of beams, 48
Bending moments, 61, 96, 126, 165,
210
Bernouli-Euler beam, 48
Bessel functions, 102, 148
Bessels equation, 88
Body
elastic, 29
piezoelectric, 36
Boundary conditions, 32, 63, 127, 165,
211, 253, 312, 349, 381
Boundary layer, 175, 219

Characteristic
length, 72, 134, 175, 177, 219, 221,
255, 314, 383
radius of curvature, 68, 170
radius of curvatures and torsion,
129, 213
Characteristic scale, 46
Characteristics, 239
Charge, 39, 193, 196, 202, 205, 233,
236
Christoffel symbols, 22, 118
of a surface, 28
Clamped edge, 61, 63, 88, 92, 111,
126, 127, 257, 351
Co-ordinates
cartesian, 18, 80, 147
curvilinear, 19
cylindrical, 23, 101, 148
elliptical, 331
of a rod, 128
of a shell, 66
polar, 87, 285
spherical, 23, 118
Co-vector, 20
Compatibility, 58, 66
Completeness, 35
Constants
dielectric, 38
elastic, 37
piezoelectric, 38
Constitutive equations

421

422
for elastic rods, 126
for elastic shells, 65, 252
for piezoelectric rods, 210
for piezoelectric shells, 168, 349
Constraint, 33
Coupling, 37, 212, 232, 234, 240
Covariant derivative, 28
Cross section
centrally symmetric, 226
circular, 144, 330, 379
elliptical, 140, 223, 331
rectangular, 142, 227, 229, 331
Cross section problems, 138, 220, 222,
327, 393
Curvature
Gaussian, 27
mean, 27
Curve
binormal to, 24
curvature of, 24
moving triad of, 24
normal to, 24
tangent to, 24
torsion of, 25

INDEX

3-D, 38, 170, 214, 350, 382


longitudinal, 172, 215
transverse, 172, 215
Electric field, 37, 171, 214
1-D, 209
2-D, 163
Electric induction, 37
1-D, 210
2-D, 165
Electric potential, 36
1-D, 208
2-D, 162
Electroded faces, 162, 167, 177, 192,
200, 347
Electroded side, 208, 211, 221, 225
Electrodes, 39, 161
Electroelastic state, 179
Elongation, 122, 208
Energy
1-D kinetic, 125, 209, 310
1-D strain, 125, 310
2-D kinetic, 60, 163, 250
2-D strain, 60, 64, 250
3-D internal, 40
3-D kinetic, 32, 40, 68, 130, 170,
214, 350, 382
Dispersion, 51, 77, 98, 105, 106, 145,
3-D strain, 31, 68, 130
146, 279, 284, 299, 334, 337
balance, 35, 44
non-linear, 54
longitudinal, 70, 131
Displacement, 29, 36
shear, 131
Donnel-Mushtari operator, 97
transverse, 70, 131
Dual problem, 139, 227, 230
Equation
Edge mode, 341
of electrostatics, 37, 180, 184
Edge resonance, 290
Equations
Eigenfunction, 34
of equilibrium, 184
Eigenvalue, 33
Equations of motion
Eigenvalue problem, 32
for elastic rods, 126, 311
Elastic enthalpy, 41
for elastic shells, 64, 253
Electric enthalpy
for piezoelectric rods, 211, 380
1-D, 209, 212, 225
for piezoelectric shells, 165, 348
2-D, 163, 168
Error estimate, 181

INDEX
Euclidean point space, 17
Euler equation, 52, 224
Extension
measures of, 58, 95, 116, 348
Fast variable, 52
Field
kinematically admissible, 183
statically admissible, 184
First approximation, 46
Fixed edge, 61, 63, 111, 127
Force, 30
Free edge, 61, 63, 89, 93, 111, 127,
198, 211, 253, 312
Frenet formulae, 25
Frequency, 32, 51, 77
antiresonant, 193, 202, 205, 232,
233, 237
cut-off, 99, 256, 300, 316, 352, 384
resonant, 193, 202, 205, 232, 233,
237
Frequency equation, 88, 90, 92, 93,
112, 120, 156, 158, 159, 192,
193, 196, 199, 202, 205, 232,
233, 237, 286, 288, 340, 342,
345, 369, 375, 400
Frequency spectra
of plates, 87, 191, 284, 367
of rods, 153, 231, 338, 398
of shells, 107, 115, 197, 304
Gauge invariance, 101
Gauss formula, 29
Gauss theorem, 23
Gibbs function, 41
Greek indices, 25
Group velocity, 78, 275
Hamiltons principle, 32, 40
for elastic rods, 125, 309
for elastic shells, 60, 61, 249
for piezoelectric rods, 210, 379

423
for piezoelectric shells, 164, 347
Helmholtzs decomposition, 91, 101,
147
Hookes law, 30
Hyperbolicity, 35, 271
Impact interval, 242
Initial conditions, 32, 39, 63, 127, 254,
312, 349, 381
Internal degrees of freedom, 249, 310,
379
Jacobian, 19
Jump conditions, 168, 212, 238
Kronecker delta, 20
Lagrangian, 45, 314
average, 73, 136, 176, 179, 261,
318, 355, 385
Lame constants, 31
Laplaces equation, 141
Legendre transformation, 40, 139, 189,
227
Legendres functions, 119
Legendres polynomials, 119
Length element, 28
Levi-Civita tensor, 101
Longitudinal impact, 238
Lower bound, 228
Mass density, 30
Mathieus equation, 332
Membrane stresses, 61, 96, 165
Metrics, 21, 27, 118
of a rod, 129
of a shell, 67
Minimizer, 49
Mirror plane, 173, 217
Mixed faces, 167, 194
Moments of inertia, 137, 220
Neumann problem, 139

424
Newtons rule, 47, 85
Normalization, 255, 352, 383
Null Lagrangian, 268, 324, 392
Orthogonality, 34, 261
Permutation symbols, 122
Phase, 51
Phase velocity, 77
Plane wave, 35
Plate, 76
circular, 87, 191, 367
elastic, 57
piezoelectric, 169
Pochhammer equation, 151, 335
Poissons equation, 140
Poissons ratio, 36
Prager-Synges identity, 181
Principal curvatures, 27
Principal terms, 49, 69, 70
Rayleigh velocity, 85
Rayleighs formula, 36
Reciprocal basis, 21
Reduction, 272
Resonance, 43
Rod
central line of, 121, 128, 207
cross section of, 121, 207, 314
curvatures of, 122
diameter, 129
displacements, 122, 208
elastic, 121, 309, 314
kinematics, 122, 208
piezoceramic, 217, 379
piezoelectric, 207
torsion of, 122
Rotation, 123, 208
Rotation axis, 173, 217
Scalar product, 17
Self-adjointness, 34

INDEX
Separation of variables, 87, 91
Series
asymptotic, 46
Shear diaphragms, 113, 307
Shell
cylindrical, 94, 107, 197, 295, 304,
372
displacements, 58, 162
elastic, 57, 249
face surface, 161
kinematics, 58, 162
middle surface of, 57, 161
piezoceramic, 174, 347
piezoelectric, 161
spherical, 66, 115
thickness, 57, 161
thickness parameter, 97
Shifter, 66
Short-wave extrapolation, 265, 320,
358, 388
Slow variables, 52
Small parameter, 45
Small terms, 45, 71, 73, 175
Spherical harmonics, 119
Spiral, 29
Stationary point, 33, 45
Strain, 29, 37, 69, 130, 171, 214
Stress, 30
Summation convention, 18
Superposition, 265
Surface
basicforms of, 26, 27, 95
co-ordinates, 25
divergence theorem, 28
moving triad of, 26
normal to, 26
tensors, 27
Symmetrization, 30
Symmetry properties, 31, 38
Tangential polarization, 189, 379

INDEX
Tension, 126, 210
Tensor, 20
contravariant, 20
covariant, 20
field, 21, 22
lowering index of, 21
product, 20
raising index of, 21
transpose of, 29
Tensor field
divergence of, 23
gradient of, 22
Tensors
contraction of, 22
Thickness polarization, 187, 191, 197,
234, 347
Timoshenko beam, 49
Torsional rigidity, 143
Traction, 32
Transverse isotropy, 173, 217
Triad, 122, 207
Twist
measure of, 124, 208
Twisting moment, 126, 210

425
Velocity, 30, 37, 60, 124, 209
Vibrations
axial-radial, 110
flexural, 87, 155, 284, 306, 310,
343
free, 32
longitudinal, 91, 191, 231, 287,
306, 309, 338
low-frequency, 70, 132, 171, 214
plane, 157, 234
radial-tangential, 118
tangential, 118
thickness-shear, 256, 315, 352, 383
thickness-stretch, 256, 315, 352,
383
torsional, 109, 154
Voltage, 162, 192, 195, 232, 235
Volume element, 23
of a shell, 67

Wave equation, 101


Wave number, 51, 77
Waves
amplitude of, 51
axial-radial, 99, 106, 303
dilatational, 36
Unelectroded faces, 161, 163, 174, 191,
flexural, 76, 146, 274, 303, 337
197
harmonic, 51, 77, 145, 146
Unelectroded side, 207, 209, 218, 222,
longitudinal, 79, 146, 151, 276,
379
334
Upper bound, 228
reflection of, 291
shear, 36
Variation, 33
standing, 115
Variational principle
torsional, 99, 105, 146, 150, 302
dual, 41
Wedge product, 27
Variational-asymptotic
Weingartens formula, 29
method, 44
procedure, 46, 70, 132, 174, 218, Whithams method, 51
257, 314, 351, 382
Youngs modulus, 36
Vector, 18
field, 18
product, 17

You might also like