You are on page 1of 23

j . Sound Vib.

(1964) i, 88-i lo

STRESS

WAVES

IN S O L I D S

H. KOLSKY
Brown University, Providence, Rhode Island, U.S.A.
(Received 19June I963)
A large and growing number of original papers on both the experimental and the
theoretical aspects of stress wave propagation is appearing in the scientific literature, and
two international conferences solely concerned with the subject have been held during the
last five years. The purpose of this paper is to review recent experimental and theoretical
advances in the propagation of deformation waves of arbitrary shape through elastic and
anelastic solids, and also to attempt to outline the problems on which present efforts are
being directed and to predict probable lines of future development.
INTRODUCTION
The subject of wave propagation in elastic solids has a long and distinguished history.
The early theoretical work is associated with names such as Navier, Poisson, Stokes,
Rayleigh and Kelvin. At the beginning of this century, when the concept of an elastic
ether was finally abandoned, interest in the whole field died down; and with the exception
of seismological studies, which were largely concerned with the application of earlier
theory to rather complex physical situations, very little new work was done. During the
last two or three decades there has been a remarkable revival of interest in the field. A
large and growing number of original papers on both the experimental and the theoretical
aspects of the subject is appearing in the scientific literature, and two international conferences solely concerned with stress-wave propagation have been held during the last
five years (l-Z). At the end of this paper, references are given to a number of books, articles
and reviews (3-17) which have appeared in the last decade and which are concerned with
various aspects of the subject.
There are a number of reasons for this revival of interest, of which the most important
are: first, the availability of experimental apparatus with which stress waves can be
produced and detected; second, the practical engineering need for data on the behaviour
of materials at very high rates of loading when stress waves are inevitably generated;
third, the growing interest in the structure of solids, where high frequency stress waves
provide a powerful experimental tool for investigating the microscopic processes which
take place in a solid when it is deformed; and last (and unfortunately, not least) the
universal increase in armament research where the investigation of the response of
structures to large forces maintained for very short times plays such a prominent part.
One class of investigation in which there has been considerable and continuing interest
is that of the propagation of pulses of high frequency sinusoidal oscillations of very small
amplitude. This has in fact become a subject of its own, known as Ultrasonics, and in the
present article it is not proposed to consider in any detail either the many and important
results on the structure of solids and fluids which have been achieved by the use of this
technique, or the use of such pulse techniques for flaw detection in engineering components. T h e purpose of this paper is instead to review the recent experimental and
thcoretical advances in the propagation of waves of larger amplitude and arbitrary shape
through elastic and anelastic solids, to attempt to outline the problems on which present
efforts are being directed and to predict probable lines of future development.
88

STRESS WAVES IN SOLIDS

89

From a theoretical standpoint the subject falls naturally into two parts which correspond
to whether or not the medium through which the stress wave is being propagated can be
considered to obey Hooke's Law. The experimental techniques used are generally similar
whether elastic or anelastic solids are being investigated, but the purpose of carrying out
experiments is often quite different in the two types of investigation. Thus, in principle,
any elastic wave problem can be treated theoretically when the values of the elastic
constants of the medium are known. In practice very few such problems indeed can be
solved, and experiments are carried out either to test the validity of approximate theories
or to obtain information about the behaviour of systems which are far too complicated for
theoretical studies even to be attempted. Such experimental investigations are then the
dynamic counterpart of the experimental stress analysis of engineering structures.
In the second type of investigation, where the solid through which the waves are being
propagated cannot be assumed to obey Hooke's Law, experiment plays a different role
as in nearly every case an "exact" theoretical treatment is of very limited validity since it
depends on a mathematical model of the anelastic behaviour of the solid. Now, for many
materials Hooke's Law is an extremely close approximation to the actual mechanical
behaviour of the solid when the deformations are sufficiently small, and the law is of the
same form, i.e. a linear one, for all elastic solids. In contrast, the various mathematical
representations of anelastic behaviour which have been used are approximate ones, which
describe specific solids, are of very limited validity, and often would appear to have no
physical validity at all. Thus experimental investigations are generally the only possible
method of determining the real behaviour of the materials, and theory can, at best, act
only as a general guide to the type of behaviour which may be expected.
Departures from Hooke's Law may manifest themselves in two ways: first, the stressstrain relation may depart from linearity and also be different for loading and unloading;
such behaviour is characteristic of metals strained" plastically" beyond their proportional
limit; second, the stress-strain relation may depend on the rate of straining so that the
material behaves in a manner analogous to that of a viscous fluid; such materials are called
viscoelastic, and viscoelastic effects are most clearly seen in high polymers such as rubber
and the many different plastics. As will be described later, many such viscoelastic materials
are "linear" in the sense that although they do not obey Hooke's Law, their behaviour
can be described by linear differential equations involving the stress, the strain, and their
derivatives with respect to time. This division into non-linear behaviour, which does not
depend on strain-rate, and linear viscoelastic behaviour, which does not depend on
amplitude, is convenient although often somewhat artificial. Under some conditions
many anelastic materials behave both in a non-linear and in a time-dependent manner so
that the stress-strain relation, or the constitutive equation, as it is often called, must be
expressed in the form of a non-linear differential equation involving stress, strain and time
and is often of considerable complexity.
However, since on the one hand the mechanical behaviour of metals is generally not
highly rate-dependent, and on the other hand most unfilled polymers and rubbers are
linearly viscoelastic for small deformations, theoretical studies in which the non-linearity
and the time-dependence are dealt with separately have considerable practical justification.
Waves travelling in a rate-dependent solid are termed viscoelastic waves, whilst waves
travelling in a solid which shows the phenomenon of yield, so that the stress-strain curve
is concave towards the strain axis, are termed plastic waves. When the stress-strain curve
is non-linear but curves towards the stress axis, i.e. becomes "stiffer" for large stresses
than for smaller ones, the nature of wave propagation is radically different from that of
either viscoelastic waves or of plastic waves.
When a mechanical pulse is propagated through such a medium, the velocity of propaga-

90

ii. KOLSKY

tion is greatcst h)r those parts of tile wave whosc amplitudes arc greatcst, and conscquently
the front of the disturbance becomes stccper and steeper :Is the pulse progresses through
thc medium. The final shape of the wave is then determincd by dissipative processes
which take place because of the intensc stress gradients produced in the wave front, and
often also depends on the microscopic structure of the medium. Such waves, which are
called shock waves, of course also occur in fluids.
In discussing and describing recent work in thc field of stress waves in this articlc, the
four types of problems outlincd above will be considered separately, viz: elastic waves,
viscoclastic waves, plastic waves and shock wavcs.
ELASTIC WAVFS
Almost the only problems for which a complete wave solution is available are the
following:
(a) TIlE INFINITE ISOTROPIC ELASTIC SOLID

Only two types of wave can propagate, namely, dilatational or P waves which travel with
the velocity ct = ~/[(k + ~G)/p] and distortional or S waves which travel with the velocity
c2 = x/(G/p), k is here the bulk modulus, G the shear modulus and p the density.
(b) TIlE SEMI-INFINITESOLID
In addition to the P and S waves, Rayleigh surface waves, which travel with a velocity
c, a little lower than that of S waves, are propagated parallel to the free surface. The ratio
G/c2 depends on the value of Poisson's ratio for the material and for real materials must lie
between o.874 and o.955.
(C) TIlE INFINITE PLATE

Lamb (18) has obtained complete solutions for two types of plane waves which he has
termed symmetrical and antisymmetrical waves. These are longitudinal and flexural in
character respectively, and unlike the other waves considered above, these waves are
dispersive, i.e. their velocity of propagation depends on their wavelength A, or rather on
A/d, where d is the thickness of the plate.
When A/d is very much greater than unity, symmetrical waves travel with the velocity
.~v/{E/p(l -v2)}, where E is Young's modulus and v is Poisson's ratio. On the other hand,
antisymmetrical, or flexural, waves, travel with vanishingly small velocities as A/d tends
to infinity. When A/d is very much smaller than unity, both symmetrical and antisymmetrical waves travel with velocities which asymptotically approach % the velocity of
Rayleigh surface waves.
(d) CYLINDRICAL BARS OF INFINITE LENGTH
Three types of wave can be propagated along such bars, namely: extensional waves,
torsional waves and flexural waves. Although the solution of this problem was first
produced by Poehhammer (i9) and Chree (2o) during the last century, it is only during
the last twenty years or so that numerical calculations based on this solution have been
made, or that the experimental and theoretical problems associated with the propagation
have been considered in any sort of detail. The theory shows that, in general, all these types
of waves are dispersive, and in addition that waves of any one wavelength may be propagated in a number of different "modes" along the cylinder. For longitudinal and torsional
waves these modes are associated with the presence of nodal cylinders where one component of the motion vanishes. For longitudinal waves the nodal cylinders correspond to

STRESS WAVES IN SOLIDS

9x

the absence of motion in the radial direction, whereas for torsional waves they correspond
to no motion at all; the annuli on opposite sides of the nodal cylinder rotate in opposite
directions.
For the fundamental mode extensional waves travel with velocities which depend on
Poisson's ratio v and on the value of A/a, where a is the radius of the cylinder andA is the
wavelength. When A/a >>I, the waves travel with the velocity X/(E/p)= Co. When
A/a < I, they travel with the Rayleigh surface wave velocity G. For intermediate values of
A/a they travel at phase velocities between Co and c,.
Flexural waves in the fundamental mode travel at vanishingly small velocities as A/a
tends to infinity and at G asA/a tends to zero. Torsional waves travel at constant velocity
c2 = "v/(G/p) for all wavelengths so long as the propagation is in the fundamental mode.
Higher modes, whoever, show dispersion.
The theory of wave propagation in solid circular cylinders was extended by Ghosh (2 x)
to consider wave propagation along hollow cylinders. This problem was also considered
by Glebe and Blechschmidt (22), who obtained an approximate solution which is inaccurate at high frequencies. Hermann and Mirsky (23) and Greenspan (24) have evaluated
the dispersion curves resulting from Ghosh's solution. The problem is of interest in that
for very long wavelengths the hollow cylinder behaves as a rod, i.e. there is no dispersion,
and the velocity of propagation is ~/(E/p); whereas, for very short wavelengths it behaves
as a plate, and when the wavelength becomes short compared with the wall thickness of
the cylinder, the velocity of propagation approaches the Rayleigh surface wave velocity c,.
The extreme complexity of exact solutions in the theory of elastic wave propagation has
led to a number of "approximate" theories being developed. Instead of solving the elastic
equations for the prescribed boundary conditions these theories use physical considerations to determine the likely nature of the motion. The earliest of these was that due to
Rayleigh and Love, who modified the simple theory of propagation of extensional waves
along a rod to allow for the radial motion associated with the lateral expansions and
contractions produced by the finite value of Poisson's ratio. Other notable approximate
treatments are those due to Timoshenko (25) for flexural waves in beams where rotary
inertia and shear deformation are taken into account, and the theory due to Mindlin and
Hermann (26) for extensional waves in rods where allowance is made not only for radial
motion but also for radial shear. This theory has recently been further developed by
Mindlin and McNiven (27). An excellent review of the many other recent approximate
and "exact" approaches to elastic wave propagation is given in the article by Miklowitz
(i6).
Although considerable theoretical work has been carried out on elastic wave propagation, many important questions, even for the very simple geometries capable of analytical
treatment, remain largely unanswered. Thus, in the problem of wave propagation along
an infinite elastic cylinder, the Pochhammer-Chree solution shows that propagation can
occur in an infinite number of modes. How the energy of an arbitrary disturbance is
divided between these separate modes cannot, in general, be determined theoretically.
The theory for an infinite cylinder is based on the boundary condition that at the free
cylindrical surface both the normal stress and the shear stress shall vanish. For a finite
cylinder, however, we have the additional boundary conditions that the ends of the cylinder
shall be free from stress. As shown in Love (28) the Pochhammer-Chree solutions do not
correspond to such end conditions, and no exact solution for wave propagation in an
elastic cylinder of finite length has so far been obtained.
Another problem to which it is difficult to find a theoretical answer is the signalvelocity
of a disturbance along a cylinder of finite length. In an unbounded elastic medium waves
can travel with one of two velocities, the dilatational velocity ct and the distortional velocity

92

it. KOLSKV

C2. Consequently, when the end of a cylindrical elastic rod is struck, some energy" would
be expected to travel along it at the velocity c~ since a dilatational wave will travel out
from the point of impact, and a portion of the wavefr,nt (albeit a very small portion if the
length of the cylinder is very much greater than its radius) will reach the opposite end
without being reflected at the cylindrical boundary. Now the Pochhalnmer theory as
normally presented shows that the group velocity of longitudinal waves is never greater
than Co = ~/(E/p); and since co is considerably less than Q, the theory clearly, doesnot
take into account such precursor waves. Adem (29) would appear to have been the first to
point out that in addition to the progressive waves treated by Pochhammer, the equation
can be satisfied by solutions inw)lving complex wave numbers. Such solutions correspond
to waves which are attenuated exponentially with distance of travel. Although such
solutions will have no relevance for an infinite rod, they mav well make it possible to satisfy,
the condition of zero end stresses in a finite rod, and to account for the propagation of
disturbances at velocities greater than co.
Experimental work on testing the predictions of the Pochhammer-Chree theory and
seeing to what extent modes higher than the first are excited in practice, has been carried
out by' Davies (3 o) and subsequently by a number of other workers (31-33). In these
investigations pulses produced by impacts or by the detonation of explosives at the end
120
I O.
lOOt

.~-..

....

9o!
1
1
1

70!-

~ 5O

40

Z o;/ /
,oLY:
0

_
8t-~ _12;_ i 16:- ~ 20:-~24~

218- ' . . .3'2


....

36

Time (microsec)

F i g u r e I. C o m p a r i s o n betnveen o b s e r v e d d i s p l a c e m e n t - t i m e curve in steel rod and ordinates


o b t a i n e d by F o u r i e r s y n t h e s i s . E x p e r i m e n t a l c u r v e ; :- calculated o r d i n a t e s ; - - u n d i s t o r t e d
pulse.

of elastic cylinders have been observed after they have travelled various distances. Various
types of pulse-shape, and cylinders covering a range of diameters have been used, and the
observations have all shown that, in practice, the fundamental mode is the one normally
excited and the predictions of pulse shape based on this mode are found to be in excellent
agreement with the observed results. Figure I shows a comparison, carried out by Hsieh
and the author (33), between the predicted and observed pulse shape in an experiment in
which a small lead azide charge was detonated at one end of a steel rod 0-2 5 in. in diameter
and 9 in. long; the observations of displacement were made at the opposite end of the rod.
On the same figure, the displacement-time curve corresponding to the initial pulse is
shown. It may be seen that the theory predicts, and experiment bears out, a considerable
lengthening of the pulse during propagation and the appearance of a series of high frequency
oscillations at the end of the pulse. These result from the lower group velocities of high

STRESS WAVES IN SOLIDS

93

frequency components in the fundamental mode. No evidence of propagation in the higher


modes was found in the experiments listed above. Some workers (34-36 ) claim to have
found definite evidence for higher mode propagation, particularly at points close to the
source of the stress pulse. This is where such effects are to be expected, since the group
velocities of low frequency waves in the second and higher modes are very much smaller
than for the fundamental mode.
Experiments designed to measure the signal velocity for extensional waves in cylindrical
bars have been carried out by Hughes, Pondrom and Mims (38), by Kolsky (39) and by
Miklowitz and Nisewanger (34). These experiments have all shown that a small part of
the pulse does travel with the dilatational velocity cl as would be expected on physical
grounds. For cylinders where the ratio of length to radius is reasonably large (of the order
of xo or more), however, the amplitudes of the precursor waves are extremely small.
A considerable amount of experimental work has also been carried out on the propagation of flexural pulses along elastic bars (4o, 4Q. The results were found to be in good
agreement with the predictions of the exact Pochhammer theory and of the Timoshenko
(25) approximate treatment; the latter corresponds very closely to the exact theory for
the fundamental mode. Ripperger and Abramson (4 I) claim to have found a signal velocity
of cl for the propagation of flexural waves although, here again, the amplitude of the
precursor was found to be very small indeed.
Very little experimental work appears to have been carried out on torsional wave
propagation in rods. Experimental work of Owen and Davies (42) and of Krafft (43) shows
that no dispersion occurs and presumably all the energy travels in the fundamental mode.
In these experiments Owen and Davies (42) produced the torsional waves by transverse
impact of bullets on the side of the bar whilst Krafft (43) used axial impact with a rifled
bullet; torsional waves were produced as a result of the frictional resistance of the bar
to the rotational motion of the bullet.
The physical reason for the extreme complexity of elastic wave propagation in a
bounded elastic solid as compared with the relative simplicity of propagation in an
unbounded medium, is associated with the nature of the reflection of an elastic wave at a
free boundary. It may be shown--see for example (3)--that when a dilatational (P) wave
is obliquely incident on a free surface, the conditions for vanishing normal and shear
stresses at the boundary require that both a dilatational and a distortional (S) wave be
reflected. The direction in which these two reflected waves travel obeys Snell's Law,
i.e. the reflected P wave makes an angle equal to the angle of incidence; whereas the
"refracted" S wave travels in a direction such that the ratio of the sines of the angles to
the normal is equal to the ratio of the velocities of propagation of S and P waves in the
solid, i.e. the "refractive index" is cl/c2. The amplitudes of the reflected P wave and
refracted S wave depend on the angle of incidence and on the value of Poisson's ratio of
the elastic medium. Similarly, when an S wave is incident on a free boundary, a reflected
S wave and a refracted P wave are generated.
If a disturbance is set up at a point in a bounded elastic solid, in general both a P and
an S wave travel out from the point. As each of these waves reaches a free boundary, it
is reflected and produces two waves; and these, in turn, travel through the solid and are
reflected at its free boundaries to produce more waves. It may thus be seen that in any
bounded solid an extremely complicated wave pattern is generated even after very few
reflections have taken place.
The simple theory of reflection at a free boundary is adequate at all angles of incidence
less than ~r/2, although for wave trains of finite breadth, diffraction effects must be taken
into account. For glancing incidence of a P wave, however, the theory breaks down, since
the simple treatment predicts a reflected P wave, travelling parallel to the surface, which

94

11. KOLSKY

is of equal amplitude and opposite sign to the incident wave. Thus all motion in the
medium ceases, and we have the trivial solution of zero stress and zero particle velocity
everwvhere. The theory of reflection is based on the concept of a train of waves of infinite
breadth and of infinite duration, and equilibrium between the incident and reflectcd
waves is assumed to have been set up. The practical problem of reflection at glancing
incidence, however, is concerned with the transient conditions which obtain when a
dilatational wave reaches a boundary parallel to its direction of travel, and this problem
has been considered by Sauter (44), Goodier and Bishop (45) and Rocslcr (46). Sauter's
analysis is the most thorough and takes into account diffraction effects, which depend
on the wavelength, as well as on the geometrical conditions of reflection. The expressions
he derives are, however, extremely complicated and difficult to apply to real physical
situations. Goodier and Bishop (45) and Roesler (46) have treated the problem by using
a limiting process for the reflection of waves as the angle of incidence approaches 7r/2.
These analyses show that an S wave is generated, and the amplitude of the incident P
wave is attenuated as it progresses along the boundary.
Schardin (47) and Christie (48) have shown experimentally that such a trailing S wave
is in fact produced. The experimental technique employed by these investigators was to
propagate waves in plates of glass and other transparent elastic materials, and observe the
reflection at glancing incidence by means of a photoelastic technique; a series of highspeed photographs of the photoelastic pattern was taken while the pulse was being reflected. Christie (48 ) was able to make some quantitative measurements of the intensity
of the reflected S wave at very large angles of incidence and at glancing incidence, and he
found his experimental results were in good agreement with the numerical predictions of
the treatment developed by Roesler.
Other experimental investigations on elastic wave propagation which should be
mentioned are those carried out to simulate seismological situations. The specimens
employed have been either large blocks of concrete and granite (49-5o), where the threedimensional problems could be tackled directly, or plate specimens (49--52)- As shown by
Bishop (53), the propagation of waves in plates and two-dimensional propagation in
infinite solids are similar except for the numerical values of the elastic constants.
Most of this model seismological work has been carried out by the use of piezoelectric
or ferroelectric crystals as generators and receivers. The technique has differed from that
normally used in ultrasonics in that the crystals are pulsed by a high transient electrical
d.c. potential (> mooV) for a few microseconds, and individual pulses rather than
sinusoidal wave trains are generated. To ensure this, the crystals employed are ones whose
natural frequencies are remote from the predominant Fourier components in the voltage
pulse, so that "ringing" does not occur. Sherwood (52), in model seismological studies
with plates, used small explosive charges to generate the stress pulses, and condenserunits to detect and measure them.
It will be seen that most of the experimental work carried out on elastic waves hitherto
has been concerned with studying propagation in specimens of comparatively simple
geometrical shape, where the experimental results could be compared directly with
exact or "approximate" theoretical predictions. With increasing confidence in the
experimental techniques and in the interpretation of the observations, it should be
possible to study more complicated problems. One such problem that has been tackled
expcrimentally by Ripperger and Abramson (54) is the passage of longitudinal and
flexural elastic pulses across the discontinuity in cross-sectional area between two cylindrical bars of different diameters. This problem, which involves discontinuous boundary
conditions in stress and displacement at the junction between the two bars, cannot be
treated theorctically with any pretence of rigour. It is a good example of the type of

STRESSWAWS m SOLIDS

95

problem, the solution of which is likely to be of practical importance, but which only
experiment can hope to elucidate.
In addition to studying more complicated geometries, future work in elastic wave
propagation is likely to be concerned both theoretically and experimentally with the
diffraction phenomena. These occur when a mechanical pulse starts to be propagated in a
specimen, the lateral dimensions of which are of the same order of magnitude as the
wavelengths of the dominant components of the Fourier spectrum of the pulse. As in
optics or acoustics, a ray theory then becomes inadequate, and we enter the area which
for light is termed physical optics in contrast with geometrical optics. In the propagation
of elastic waves this generally involves the transition between conditions of plane strain
and conditions of plane stress; and when neither of these can be expected to apply, the
problem is one of considerable complexity even for specimens of very simple geometrical
shape.
VISCOELASTIC W A V E S

W h e n a specimen of any real solid is taken round a stresscycle, some of the mechanical
energy is lost as a result of dissipative mechanisms in the material. T o this extent the
concept of elasticwave propagation, as considered in the previous section isan idealization,
and the equations of motion should always include terms corresponding to dissipative
processes in the material. In practice, however, for most crystalline materials and for
many amorphous materials, such as glass, when the deformations are sufficientlysmall
the losses due to internal friction are extremely small; and the experimental results on
wave propagation are found to be in excellent agreement with the assumption of linear
elastic behaviour.
In contrast, many of the materials which are composed of giant organic molecules and
are known as high polymers, show very large internal dissipation. With the growing
engineering importance of these materials, which include plastics and rubber, considerable
effort has been directed in recent years to the study of the dynamic behaviour of structures
composed of such materials, and also to the problems of stress wave propagation through
such viscoelastic solids. In addition to the high mechanical dissipation that these materials
exhibit, they also show a very marked dependence of the values of the elastic" constants"
on the rate of application of stress. Thus, Young's modulus for a metal like steel will be
found to have substantially the same value, whether the measurements are carried out
in a conventional testing machine, or are obtained by measuring the velocity c of an elastic
pulse along a thin wire of the material and using the relation E = pc2. A similar comparison
for a filament of natural rubber will show the value of E in the wave experiment to be
considerably greater than that found in the testing machine. Under some conditions the
"dynamic" value obtained from the wave propagation experiment may be as much as a
thousand times the "static" value obtained in the testing machine. This dependence of
elastic modulus on time of loading means that from the point of view of wave propagation,
the material is dispersive, i.e. sinusoidal waves of high frequency have higher velocities of
propagation than low frequency waves. It is also found that the internal friction results in
high frequency waves being attenuated more rapidly than waves of lower frequency. As
a result of these two effects, i.e. velocity dispersion and frequency-dependent attenuation,
the shape of a mechanical pulse as it is propagated through a viscoelastic medium changes
rapidly even in the absence of the type of boundary effects discussed in the previous
section.
As mentioned in the introduction to this article, the type of viscoelastic response on
which most theoretical and experimental effort has hitherto been expended is that termed
linear viscoelastic behaviour. A linear viscoelastic solid may be defined in a number of

96

ii. KoI.SKY

different ways which can be shown to be mathematically equivalent (55). One definition
is that it is a material for which the stress components and tile strain components are
related by linear differential equations which involve tile stress, the strain, and their
derivatives with respcct to time. Thus when we consider uniaxial deformations, so that
only one component of stress a, and one component of strain E, are relevant, the stress-strain relation may be written as
Po-=

QE

(I)

where P and Q are linear differential operators so that

i' = ao+al D + a 2 D 2 . . . .
and

Q = bo+blD+b2D 2....
D is here the operator d/dt and the a's and b's are numerical constants.
Historically, a linear viscoelastic solid was first defined by Boltzmann (56), in terms of
his principle of linear superposition of time-dependent mechanical response. In Boltzmann's treatment it is assumed that the mechanical response of a material depends on its
entire previous loading history ; and that if a specimen is subjected to a number of separate
deformations, the stresses at subsequent times will be the simple algebraic sum of those
which would have been found to be acting if the deformations had been carried out singly.
Alternatively, if the specimen is subjected to a series of stress cycles, the total resulting
deformations will be simply the sum of the individual deformations produced by the
separate stress cycles. Mathematically, the principle can be written for uniaxial stress
and strain as either
t

= .I f ( t - r)
--

dr,

(2)

ts3

w h e r e f ( t - r ) is called the stress relaxation function, and a and E are the stress and strain
respectively, or as
t

E=

f g(t-r)(do/dr)dr,

(3)

zo

where g ( t - r) is termed the creep function.


Boltzmann's treatment is physically perhaps the most satisfactory approach to linear
viscoelastic behaviour, and the stress-relaxation function can be measured experimentally
by observing how the stress varies with time at constant extension, whilst the creep
function can be measured by observing the extension at constant stress. This is in contrast
with the operator functions P a n d Q in equation (i), which have to be fitted to experimental
results, a sufficient number of a and b terms being included to give a reasonable fit.
It may be shown that where there are a finite number of terms in the operator equation,
the behaviour can be represented by a mechanical model consisting of springs which are
perfectly elastic, and viscous elements, called dashpots, for which the stress is proportional
to the time rate of change of the extension, i.e. springs which are Hookean and dashpots
which obey Newton's Law of Viscosity. One simple model of this type is the Maxwell
model, which consists of a spring in series with a dashpot. For this model P = a0 + al D
(and all the other a's are zero) and Q = bl D (and all the other b's are zero). Another simple
model is the Kelvin-Voigt model which consists of a spring connected across a dashpot.
For this model P is a constant and Q = bo+biD.
It is found (57) that real viscoelastic solids, such as rubbers and plastics, behave in a
much more complicated manner and can be described by such simple models only when

STRESSWAVESIN SOLIDS

97

a narrow spectrum of loading times is involved, i.e. when the ratio between the longest
time and the shortest time that is relevant is not more than, say, so.
The disadvantage of the integral representation as derived from the Boltzmann superposition principle is that it leads to difficult integral equations when k is applied to specific
problems. The difficulties with the use of the operator method are first, that finding the
appropriate operators involves considerable computation and curve fitting, and second,
that if a large range of times has to be covered, the P and Q operators must each contain a
considerable number of terms. This means that in trying to tackle a specific problem
one may be faced with differential equations of the fifth, sixth or higher order, and these
may prove quite intractable. There is another approach to linear viscoelastic behaviour,
namely the use of complex elastic moduli. This again mathematically is equivalent to the
methods outlined above and hence inherently involves many of the same difficulties, but
for many purposes, and particularly wave propagation, it is extremely convenient.
The computations which are required when this method is applied to specific problems,
are generally in the form of the summation of Fourier series, and electronic computer
programmes, as well as numerical methods which use hand machines, are readily available
for carrying out computations of this type.
When a linear viscoelastic solid is subjected to a stress o that varies sinusoidaIIy with
time at an angular frequency p, the strain produced will vary sinusoidally with time at
the same frequencyp; but in general there will be a phase lag ~ between the strain and the
stress.
At any one frequency the phase lag 8, and the ratio between the stress amplitude o0,
and the strain amplitude 0, will be constants independent of the absolute values of the
amplitude. Using complex notation we may write
= o0 exp (ipt),
and
so that
or

e = 0 exp (ipt- i8),


o/ = (o0/e0) exp (i~) = E1 + iE2,
a = (Et + iE2).

(4)

Here E, and E2 are termed the real and imaginaryparts of the complex modulus E. It
may be seen that gz/g, = tan ~ and that o0/0 = ~/(E~ + E~); this ratio is often denoted by
E*.
Now E* and tan8 are quantities which can be obtained directly from vibration experiments and, if their values are known over a sufficiently wide range of frequencies, the
mechanical response of the solid to any uniaxial mechanical deformation can be calculated
by Fourier synthesis.
We now return to the problem of wave propagation in such a linear viscoelastic solid.
If we consider the simplest type of propagation, namely a longitudinal wave travelling in
the x direction along a thin rod, where the effects of lateral inertia may be neglected,
Newton's second law of motion gives the relation

ao/ax

;(a2u/at2),

(5)

where p is the density of the rod and u is the displacement in the x direction. Equation (5)
will hold whatever the stress-strain relation may be.
For a stress which is varying sinusoidally with time in a linear viscoelastic solid, we
may use relation (4), so that (5) becomes

(Et + iE2) (0 2 u/ax 2) = p(a 2 u/Ot2),


since

= au/ax.

(6)

98

H. KOLSKY

The solution of (6) for a sinusoidal displacement u ~ u . e x p ( i p t ) at the origin x = o


and a wave travelling in the direction of increasing ,v is
u = u,,exp[-oa'+ip(t

- x c)],

where

,"-=

(E*;>)'Zsec(8/2), !

and

(p/c) tan

:=

[(E2/E , = tan8

(a/z);

and

E*=

(7)
(s)

v'(E~+E~)].

Thus, if we know the value of E I and E2, or of E* and tall 8 at any frequency p, equations
(7) and (8) give us the phase velocity c, with which such a sinusoidal wave is propagated,
and the coefficient ~, with which the amplitude of the wave is attenuated with increasing
x as the wave progresses along the rod.
Now since the principle of superposition is applicable, any disturbance travelling along
the rod may be considered as the sum of a spectrum of Fourier components, and the
propagation of the pulse can be represented as a Fourier integral. Thus we can express
the displacement pulse produced at the origin, by the relation
u = j Ap exp (ipt) dp,
O

where Ap is in general complex; then at any distance x along the rod we have
u = j A, exp [ - c~x + ip(t - x/c)] dp.

(9)

and c are here of course functions o f p as given by equation (8).


In order to solve the specific problem of a mechanical pulse propagated along a viscoelastic rod, where the shape of the pulse is known at one end of the rod, we must first make
a Fourier analysis of the initial pulse. The cosine and sine terms in this analysis give the
real and imaginary parts respectively of Ap in equation (9). We must also know the values
of c and ~, or E* and tan8, over the frequency range covered by the relevant Fourier
components (for a reasonably smooth pulse fifty Fourier components give an extremely
close fit, and the viscoelastic properties need to be known over a range of frequencies of
this order). The Fourier integral in equation (9) can then be approximated by a Fourier
sum, and the shape of the pulse at any distance x along the rod can be computed.
The author (58) has carried out a number of such computations, and the results of these
were compared with experimental measurements of pulse shapes in viscoelastic rods.
Figure 2 shows such a comparison between theory and experiment. In this experiment the
pulse was produced by a small explosive charge detonated at one end of a polyethylene
rod 6o cm long and 1.25 cm in diameter. The shape of the pulse arriving at the other end
of the rod was measured with a condenser unit. It may be seen that the shape computed
from the response of the material to stresses which vary sinusoidally with time agrees well
with that observed in this experiment. In the experiment a certain amount of geometrical
dispersion was produced as a result of lateral inertia effects in the bar. These were discussed
in the previous section. It was found, however, that this could be allowed for by carrying
out the computations for a value of x slightly greater than the actual length of the bar.
In the treatment described above, the experimentally measured values of c and c~, or
of E* and tan3, have to be inserted in the computation, so that the measured properties
of the specific solid under investigation have to be used. Now, for a large class of viscoelastic solids over a considerable range of frequencies and for a range of temperatures

99

STRESS W A V E S I N SOLIDS

thc value of tan3 isfound to be comparatively independent of frequency. This isadmittedly


only an approximation, the validityof which is more justifiedfor some solids than for
others. If, however, we assume that tan3 is in fact independent of frequency and is small
compared with unity, it may be shown (I I, 5 8) that the modulus E* at frequency p can
be approximated to by the relation
E'=

log,(p/po)],

E~[z4 2ta__n3zr

where Eg is the value of E* at some reference frequency P0. From equation (8) we then
have

,,o,
and
~ -- P [I - tan
~r 3 lg* (P/P)] tan ~ '
and, since tan3 < I,
"~"

Kp,

(II)

where K = tan3/2Co.

950

900

850

'+l
o

-6

IrO
Log(#po)

I
20

30

po= 2 ~ x 103 cydes/sec

Figure 2. Velocity and attenuation for polyethylene rods. Curve I: Phase velocity (m/see);
Curve 2 : Log ct.
Thus to a first approximation we should expect to find that c plotted against logp should
give a straight line of gradient (tanS)fir, and log~ plotted against logp should give a straight
line of gradient unity. Figure 2 shows such plots for the experimental results obtained
by Hillier (59) for sinusoidal waves of different frequencies travelling along filaments of
polyethylene, and it may be seen that these results agree very well with the predictions
of the above approximations.
If we take equations (IO) and (i x) to represent the propagation constants of a class of
viscoelastic solids, it is possible to treat in non-dimensional form a number of problems
concerned with the propagation and reflection of waves in viscoelastic solids. Such
calculations have been shown to lead to predictions which are borne out by experiment

(58).

IOO

11. KOLSKY

T h e velocity dispersion produced by the viscoelastic properties of solids leads to a


nuxnber of interesting phenomena which are observed when mechanical pulses are
propagated through them. Thus, as seen above, both the velocity of propagation and the
attenuation increase with increasing frequency, so that the high-frequency components
of a pulse travel faster and are attenuated more rapidly than those of lower frequency.
As a result, a pulse, which is initially symmetrical in shape, becomes asymmetrical
(Figure 3) and at the same time becomes broader and flatter as it travels through the solid.
!

. . . . . . . . . . . . . . . . . . . . . .

c
>.

~5.-v

_,

o
-~

E
<

I00

200

300

400

50C

Line (microsec)

Figure 3. Comparison between observed pulse ill polyethylene rod and ordinates calculated by
Fourier synthesis. - - observed pulse ; o points obtained by Fourier synthesis.
Another point of interest concerns the " s h a p e " of a pulse travelling in a viscoelastic
medium. In an elastic solid the displacement in a plane wave travelling in the direction
of decreasing x can be written as
u = f(x + ct),
(z 2)
w h e r e f is an arbitrary function of the argument (x + ct).
T h e particle velocity v is then given by
v =

(I3)

~- c / ' ( x + c t ) ,

~u/~t

where the dash indicates differentiation with respect to the argument (x + ct); the strain E
is given by
=

I-O I-

~u/'?:x

. . . .

=/'(x-'-ct),
_._~_.

11

-400

-200

(z4)
,

-=.

7o

200

400

Time (mlcrosec)

I'igure 4. Comparison of the pulse shapes of displacement, particle velocity, stress and strain
in a viscoelastic rod.

STRESS WAVES IN SOLIDS

IOI

and the stress by o = E,, where E is the appropriate elastic modulus, so that
o

(IS)

= Ef'(x+ct).

From (13), (14) and (iS) it may be seen that v, c, and a are allof the same pulse shape,
and the displacement is proportional to the integral of these pulses.
For viscoelasticsolids this is no longer true since c and E are time-dependent, and the
pulse shapes will not be the same for stress,strain,and particlevelocity. From an experimental point of view this can lead to difficultiessince it is often possible to measure only
strain or particle displacement when, for example, a knowledge of the stress is required.
Furthermore, in some experimental techniques, notably the photoelastic technique, it
is not quite clear whether the pulse shape observed is that of stress or that of strain. For
elastic studies this does not matter once the detectors have been calibrated; but when
viscoelastic materials are used, and much photoelastic work is done with such materials,
it is not only the amplitude but the shape of the pulse which is suspect.
Fortunately, for smooth pulses the differences in pulse shapes are generally not too
large, as may be seen from Figure 4, which shows the results of a recent calculation
carried out by S. S. Lee and the author (6o) on pulse propagation in polyethylene. The

g
>.
o
35

;~

ioo

/~' 20~,
t-

Time (microsec)

Figure 5. Reflection of pulse (from polymethylmethacrylate rod to elastic rod).


o r incident pulse ; o, reflected pulse.

pulse shapes for v, Eand o are calculated for a prescribed distribution of the displacement u.
In the Figure the shapes are given non-dimensionally; the displacement is given as u/u ~o,
where u~o is the value of the particle displacement after the pulse has passed, and the
particle velocity, strain, and stress are given as V/Vo, c/0, and ~/o0 respectively, where the
zero subscript denotes the maximum value of each quantity.
Reflection of a pulse at the interface between two viscoelastic solids, or between a
viscoelastic and an elastic solid, is also of interest, in that the reflection coefficient depends
on the values of pc where p is the density and c is the velocity of propagation; and since
8

102

H. K O L S K Y

the latter varies with the frequency for a viscoelastic solid, the reflected and transmitted
pulses can have shapes quite different from that of thc incident pulse.
Thus, consider the reflection of a pulse travelling from a viscoelastic solid into an
elastic solid which has a value of pc independent of frequency; let this value be poCo.
Assume that the value of pc for the viscoelastic solid is less than poco for low frequencies
and greater than poco for high frequencies. When a pulse is reflected at the boundary
between two such solids, the low frequency components will be reflected in one phase
and the high frequency components in exactly the opposite phase. The results of such
reflection are shown in Figure 5, where it may be seen that the reflected pulse is S-shaped
although the incident pulse was unidirectional. 'l'his calculation was carried out bv S. S.
Lee and the author (6o) for the reflection of the pulse produced in a rod of polymethylmethacrylate (Perspex, Lucite) I ao cm in length, when a very sharp blow was applied to
one end of it. The resultant pulse was reflected at the other end, which was in contact with
an elastic rod of the appropriate Young's moduh, s. In the Figure the reflected pulse is
shown at twenty times its actual size.
All the problems of viscoelastic wave propagation that have been considered here
were concerned with uniaxial waves so that only one component of stress and one component of strain were relevant. The extension of the treatment to three dimensions is, in
theory, quite straightforward, and has an exact parallel in the development of threedimensional elasticity from the uniaxial treatment. However, instead of having two elastic
constants, e.g. the bulk and shear moduli, we must now have two time-dependent relations.
Thus, in the Bohzmann treatment, we would have two stress-relaxation functions or two
creep functions, and in the differential operator treatment we would need a P and Q for
shear and a P ' and Q' for dilatation.
In fact, it seems highly probable that volume changes in all homogeneous viscoelastic
solids are substantially elastic, and viscoelastic behaviour is confined to shear deformation.
The experimental evidence on which this conclusion is based is, however, somewhat
sparse, and more experimental work must be done before the assumption of perfectly
elastic dilatational response can be regarded as universally valid.
Wave propagation in non-linearly viscoelastic materials has hitherto received very
little attention either theoretically or experimentally. Some efforts in this direction will
be described in the next section, which discusses plastic waves, and also in the concluding
section of this article.
PLASTIC WAVFS
Plastic waves are associated with the phenomenon of yield in solids, and the theory of
their propagation has been developed chiefly in connection with the response of metal
structures to large transient stresses which produce plastic deformation. If the stresses
are applied very rapidly, inertia effects cannot bc neglected; and the problem has to be
treated in terms of wave propagation.
The first treatment of this type of problem was published by Donnell (6I) in i93o.
Donnell considered the propagation of a longitudinal disturbance along a thin rod of a
plastic strain-hardening material. The disturbance was produced by suddenly applying
a constant stress a to one end of the rod and maintaining this stress. Donnell assumed for
simplicity that the material had a bi-linear stress-strain curve, i.e. that the material was
elastic for stresses up to the yield point %, the value of Young's modulus being El, and
that for stresses greater than % the stress-strain curve had a second linear portion, the
gradient of which was EE(E2 < EO.
If a is less than the yield point %, a steep-fronted elastic pulse of constant amplitude
a0 is propagated along the rod at constant velocity co = ~/(E1/p). If, however, a > a0, there

STRESS WAVES I N SOLIDS

IO 3

will be two wave fronts. The elastic wave of amplitude a0, the front of which is travelling
at velocity Co, will be followed by a "plastic" wave of amplitude a which travels at the
lower velocity .v/(E2/p); the stress distribution along the rod is thus in the form of two
steps. The height of the first step is a0 and that of the second, a.
After Donnell's paper, the subject of plastic wave propagation appears to have received
little further attention until the beginning of the Second World War when formal solutions
of the problem for more general stress-strain relations were obtained independently in
England, the United States and the Soviet Union. (See references 3, 6, 9, [3, 62, 7I.)
The problem can be treated in either Lagrangian or Eulerian co-ordinates, and both
have been used in the various derivations of the solutions. We will here outline the
problem as treated by yon Karman (62), who used Lagrangian co-ordinates. Von Karman
considered an infinitely long wire, one end of which was suddenly given a velocity V at
time t = o, and this velocity was maintained for all t > o. As in Donnell's treatment, it is
assumed that the wire has a linear stress-strain relation up to a critical value of the stress
ao, the yield point; for stresses a > a0, the stress o is assumed to be a univalued function
of the strain E so long as a is increasing, and this function is assumed to be independent
of the rate of straining. It is also assumed that the effects of radial motion of the wire can
be neglected so that geometrical dispersion effects (of the type discussed in an earlier
section of this article) could be ignored.
The solution of this problem which satisfies the equation of motion and the prescribed
boundary conditions is that, at any time t, the distribution of strain along the wire can
be divided into three distinct regions.
(i) From the end of the wire x = o to the "plastic wave front", x = C1 t, the strain has
a constant value ~1. Here C1 = % / ( S 1 / p ) , where St is the tangent modulus of the
engineering stress-strain curve at a strain cl. It is thus the quantity do'/dE at E = ~,,
where a' is the nominal stress, i.e. the force applied divided by the cross-sectional
area of the undeformed wire.
(ii) From x = C1 t to x = Cot, the strain ~ decreases monotonically in such a way that
x = Ct at every point where pC 2 = da'/d~( = S), i.e. each value of the strain travels
down the wire with a velocity corresponding to the tangent modulus of the stressstrain curve at that value of the strain. Here Co= a/(E/p), where E is Young's
modulus for elastic deformations.
(iii) At x = cot the strain has a value c0. This is the maximum elastic strain that the
wire can maintain and corresponds to the yield stress a0.
For x > Cot, the wire is undisturbed.
The relation between the constant value of the strain ~I between the end of the wire
and the plastic wave front, and the velocity V at which the end of the wire is being extended
is
"t

--

f( \Pil

d,.

(I6)

S is the gradient of the stress-strain curve do/d~ and is a function of ~. The type of strain
distribution described above is shown by the broken curve in Figure 6.
The earliest experimental verification of the theory was carried out by Duwez and
Clark (63), who measured the distribution of permanent strain in a copper wire which
had been impulsively stretched by a falling weight. They found that the relation between
the permanent strain E1 and velocity of stretch 1/"given by equation (]6) was well borne
out by their experimental results, and the general distribution of strain along the wire
was in reasonably good agreement with the theory. A more detailed examination showed,

I0 4

II. KOLSKY

however, some serious discrepancies between theory and experiment, and ira particular
the length of region of constant strain, behind the plastic wave front, was found to be
shorter than predicted, and even within this region the strain was not quite constant.
These discrepancies were at first attributed to the complex wave patterns which arc set
up during the unloading of the wire. A careful analysis showed, however, that the unloading waves could only account in part for these discrepancies, and that the inadequacy
of the theory in explaining these results lay in the assumption that the stress-strain curve
was independent of the rate of straining, and the use of the " s t a t i c " stress-strain curve
of the wire was unjustified.
The applicability of a theory of plastic wave propagation based on a rate-independent
stress-strain curve was further placed in doubt by tile experiments of Bell (64), Sternglass
and Stuart (65), and Alter and Curtis (66). These workers investigated the propagation of
a small amplitude longitudinal disturbance along a bar which was already stressed beyond
its elastic limit. According to the rate-independent theory as outlined above, such a
disturbance should travel at the velocity v'(S/p), where S is the tangent modulus of the
stress-strain curve at the pre-stress of the bar. Since the yield point had been exceeded,
S was considerablv less than E, Young's modulus for the bar, and the superimposed
disturbance would be expected to travel at a velocity considerably less than co, the velocity
of elastic waves. Experimentally they found however, that the disturbance travelled in all
oooa,<,

. . . . . . . . . . . . . .

-.

-~

~_~ .... .

'

OC06i
;

,.

',

Rc/te ;ndependent

"~

RGte de;~ender,

0004

.;

,Y,

,C:

!5

._~

Figure 6. Strain distribution (after Malvern) (t = Ioa'4/~sec).


cases at the velocity of elastic waves Co. It would thus appear that when the additional
transient stress is applied, the material follows a different stress-strain relation from that
observed at lower rates of straining, i.e. the stress-strain curve is rate dependent.
A proper treatment of the problem would thus involve considering a medium which
is both non-linear and strain-rate dependent, i.e. a non-linear viscoelastic solid. As
mentioned earlier, the mathematical treatment of wave propagation in such a material
is extremely involved, and very few problems of this type have been treated hitherto.
One reasonably successful attempt is due to Malvern (67), who considered the propagation
of plastic waves in a wire for which the stress was a function of both the longitudinal strain
and the strain rate i. He expressed the relation for the plastic component of the strain
~2 a s

E~2 = g(m,),

(I7)

STRESS WAVES IN SOLIDS

IO 5

and for the elastic component el we have


EE, =

(,8)

Combining equations (17) and (18) we then have


at =

")'

(19)

where e = ,t + E2 is the total strain.


We also have the equation of motion
a~

a 2u

bv
= P

ax = p

(20)

since v = 0u/at (cf. equation 5), and the relation


aE

a--t

a 2U
=

a,/.)

O x a t - Ox

(21)

Equations (19) , (2o) and (21) form a system of three first-order differential equations
which can, in principle, be solved if the form of the function g(G, e) is known. For illustrative purposes Malvern used a very simple form for this function, namely
g(~,e) = K [ ~ - f ( ~ ) ]

(22)

From equation (17) it may be seen that G, =f(E) is the stress-strain relation at zero rate
of strain, i.e. the " s t a t i c " stress-strain relation, and equation (17) now implies that the
excess dynamic stress (G- a,) is proportional to the rate of straining.
In order to carry out a numerical calculation for hardened aluminium Malvern used
a hyberbolic relation, namely
f ( , ) = (2 x lO 4 - IO/,)Ib/in 2,
together with E = xoTlb/in 2 and p --- 2. 5 x 1o-41bsec2/in4.
T h e value of the yield point was taken as io41b/in 2 so that up to this value of the stress,
the deformations were elastic. K was taken to be Io6sec - 1. This gives a i o % increase in
stress over the " s t a t i c " value at a strain rate of 2oosec- 1, and this was the order of magnitude observed experimentally with metals.
Using the above values, Malvern calculated the strain distribution in a thin rod of
aluminium where one end is suddenly set into motion by a tensile impact at a velocity of
600in/sec, and this velocity is maintained constant. The distribution of strain io2. 4 microseconds after the beginning of the impact is shown in Figure 6. On the same figure the
broken line shows the distribution of the strain which would obtain if the stress--strain
relation of the material were rate-independent. The most notable difference between the
two curves is that for the rate-dependent material no region of constant strain and no
"plastic wave f r o n t " are present. It is possible that, with more sophisticated forms for
g(a, c) and f(e), a strain distribution which showed an approximation to a strain plateau
could be obtained, but Malvern suggests that any plausible form for f(c) is unlikely in
itself to produce a region of constant strain. The Malvern type of treatment accounts for
the experimental effects observed for super-imposed pulses (64, 65, 66) but would appear
to be inadequate for describing the observed nature of the strain distribution in the
vicinity of the end of the wire.
A number of investigations have been carried out to measure the extent to which the
stress-strain relation of metals depends on the rate of strain (68). At moderate rates of
strain this does not involve any particular difficulty, but at the rates of strain which are

io6

H. KOLSKY

relevant in plastic wave propagation, the cxperimcntalist tends to find himself travelling
in a vicious circle. Any apparatus designed to measure the stress-strain behaviour at
these rates of strain results in stress waves being generated because of the inertia of the
various parts of the apparatus and of the test specimen itself. Any attempt to interpret the
experimental observations requires an understanding of plastic wave propagation in a
strain-rate dependent material which in turn must be based on a knowledge of the stress,
strain-rate relation which the experiment has been designed to measure.
One method of resolving this dilemma for soft metals was devised by the author (69),
who used specimens in the form of thin discs, which were placed between two steel bars
along which stress pulses were propagated. Here, the inertia of the specimen and plastic
wave propagation in it could be neglected, and since wave propagation in the steel bars
was elastic, it could be treated analytically. This method has subsequently been used and
modified by a number of workers, and one of the most careful investigations of this type
has recently been carried out by Davies and Hunter (7o). The main drawback of the
method lies in the presence of frictional effects between the specimen and the bars ; such
effects are difficult either to eliminate or to allow for.
Another apparatus to measure stress-strain relations at very high rates of strain was
developed by L. S. Douch and the author (71). }Iere, rods of copper, aluminium, and an
aluminium alloy were fired from an air gun and impinged axially on one end of a steel bar.
16.00C - - . -

14.00C
12,00C
tb

IO.OOC
800C / /
6000't"
i
4000[

2000[

(0"02

0'04

0'06
Strain ~.

0.08

0"10

Figure 7. Stress-strain curves for aluminium. - - d y n a m i c cur~'e; - - - static curve; experimental points.

A range of velocities of impact was covered, and the values of stress were determined by
measuring the amplitudes of the elastic waves propagated along the steel bar, whilst the
corresponding plastic strains were measured by examining the specimens after impact.
Figure 7 shows a comparison between the stress-strain curve for annealed aluminium
bars obtained by this method and the "static" stress-strain curve of similar bars measured
in a conventional testing machine. It may be seen that there is a small but definite strainrate effect.
In this work (7 I) the experimental results for plastic-wave propagation in the metal
bars were compared with the predictions of the simple theory which ignored strain-rate
effects. Instead, however, of using the static stress-strain curve, as Duwez and Clark did,
the "dynamic" stress-strain curves of the t3Tc shown in Figure 7 were employed, and
it was found that ti~e agreement achieved between theory and experiment was considerably

STRESSWAVESIN SOLIDS

IO7

better. A rate-independent treatment can thus be used for describing plastic wave propagation to a reasonable degree of approximation if stress-strain relations appropriate to
the actual rates of straining are used.
Very much more work, both theoretical and experimental, needs to be done on plastic
wave propagation. Probably the most urgent problem is the extension of the theory,
continuing the work of Malvern, to include strain-rate effects. The interaction of elastic
unloading waves and plastic waves needs further study, since in any practical problem
such unloading waves must inevitably be propagated. Finally much of the work has
hitherto been concerned with one-dimensional problems, and the extension of both
theoretical and experimental studies to more problems in three-dimensional plastic wave
propagation should lead to many interesting and novel results.
SHOCK WAVES IN SOLIDS
In the previous section it was seen that plastic waves can be propagated through
materials for which the tangent modulus S decreases with increasing strain. When,
however, a disturbance is propagated through a medium for which S increases with
increasing strain, large strains travel at a higher velocity than lower strains so that the
large strains soon reach the front of the disturbance and a shock front is built up. Such
shock waves are not normally produced in metallic bars or rods since the stress-strain
curve under these conditions is normally concave towards the strain axis, i.e. S decreases
with increasing ~. Some metals, however, e.g. nickel-chrome steel, have a point of inflection in their stress-strain curves so that for sufficiently large strains shock-fronts should
be produced. This problem has been treated theoretically by White and Griffis (72) and
by Lee and Tupper (73), but no experimental work appears to have been carried out to
investigate the effect. K. W. Hillier and the author (74) have shown that for a number of
high polymers, such as rubber and nylon, the dynamic tangent modulus increases rapidly
with strain; shock fronts might therefore be expected to be formed in filaments of these
materials. The viscoelastic behaviour of such high polymers rather complicates the
problem, and experimental work will be required to investigate whether or not such
shock fronts are, in fact, generated.
All the experimental and most of the theoretical work on shock waves has hitherto been
concerned with the propagation of very intense plane compressive disturbances through
blocks or slabs of solids. The relevant modulus for elastic waves of this type is k + ]G,
where k is the bulk modulus and G is the shear modulus, and this is the value of S = do/d~
for waves of sufficiently small amplitude. For metals, as the stress amplitude exceeds the
yield point, the value of S drops as a result of plastic flow in shear, and plastic waves are
propagated. At even greater stress amplitudes, however, S begins to increase again as a
result of the increase in the bulk modulus k at large stresses. [Such increases of k with
pressure have been investigated by Bridgman (75).]
As soon as pressures for which S increases with E are reached, shock waves can be
propagated, although they may be preceded by elastic and plastic waves. At sufficiently
high pressures, however, the value of S will exceed k + -gG and the shock wave front will
overtake the elastic wave front. Pack, Evans and James (76) have studied the shock waves
propagated in blocks of metals when large explosive charges were detonated in contact
with a free surface of the block. They found that at the pressures produced by the explosive,
the shock waves in lead travelled faster than the elastic waves, whilst in steel the elastic
waves were ahead of the shock wave front. The increase of bulk modulus with increasing
hydrostatic pressure is not confined to metals, and shock waves can also be generated in
non-metallic solids. Shreffler and Deal (77) have studied shock-wave propagation in a
glass-like plastic as well as in a number of metals.

108

H. KOLSKY

Much information about the behaviour of solids at pressures higher than any that can
be achieved in other ways, has been obtained from shock wave studies, and an illuminating
discussion of such investigations together with a list of relevant references will be found
in a recent review article by Duvall (77).
CONCLUSION
It will have been seen from the foregoing sections that work on the propagation of stress
waves has been carried out by workers who have approached the problem from a number
of points of view. There is first the mathematical approach, where for more than a century
the field has provided challenging boundary value problems in applied mathematics.
Until very recently almost the entire mathematical effort in this direction has been concerned with elastic wave problems, but during the last two decades more and more
problems on plastic and viscoelastic wave propagation have been tackled. These studies
have been greatly facilitated by the advent of electronic computers which have made
numerical solutions possible for problems which would otherwise be quite intractable.
A second approach is that of the structural engineer, whose interest in stress waves
results from the necessity of designing structures which will withstand large transient
loads, where an analysis of the stress distribution which ignores inertia effects can lead to
completely erroneous predictions. The object of this type of investigation is to ensure that
critical stresses are not exceeded by any impact that the structure may have to withstand,
and the method of determining this is unlikely to be rigorous mathematically for any but
the very simplest cases, and in general will involve experimental investigations.
A third approach is that of the engineer who is studying the dynamic response of
engineering materials at very high loading rates. He may not be interested in stress waves
per se, but any attempts he makes to measure the dynamic stress strain characteristics of
his specimens result in waves being propagated in his measuring apparatus, and stresswave considerations are required to interpret his experimental results.
Last, there is the physicist who is interested in the microscopic structure of solids. Ite
uses stress waves, either in the form of ultrasonic pulses or as shock waves of high intensity,
as a tool for studying the effects of impurities in crystals, or of the motion of dislocations
in a crystal lattice, or of the effect of large pressures on the electrical properties of solids,
or simply for measuring the equation of state at very high pressures. Closer co-operation
between these many disciplines should lead to even more rapid advances in the subject.
An experimental application of stress waves which has so far not been mentioned in
this review is the use of large stress pulses, produced by high-speed impacts or the
detonation of explosive charges, to the study of brittle fracture. An account of these
investigations will be found in the book by Rinehart and Pearson (4) and in an article by
the author (78). One advantage that this technique offers is that stress distributions can
be set up which could not be achieved in any other way. For example, the interior of a
specimen can be momentarily in tension while all its boundaries are stress-free. Another
advantage of the method is that the stress is maintained for so short a time that fractures
cannot grow and traverse the specimen, hence a vast number of fracture nuclei can often
afterwards be seen in the specimen, all of which started to grow independently duringthe
stress-wave loading.
In the whole field of stress-wave studies, perhaps the most important and the most
challenging problem that faces us at present is the treatment of stress-wave propagation
in non-linear, time-dependent materials. This problem has arisen in two ways; on the
one hand, the strain-rate dependence of metals requires such a treatment, and on the
other, non-linear behaviour in viscoelastic solids such as rubbers and polymers must also

STRESSWAVESIN SOLIDS

I09

be treated by such a theory. T o some extent the emphasis in the two problems has been
different, in that strain-rate effects in metals are generally small whilst the non-linearity
is large; whereas, with viscoelastic solids it is often the non-linearity which is small,
whilst the rate-dependence is the dominant feature. T h e experimental and the theoretical
difficulties are however, very similar, and any advances made in the solution of one
problem are likely to help in the solution of the other. An investigation by S. R. Bodner
and the author (79) on stress-wave propagation in lead showed that this material had to
be considered as a non-linear viscoelastic solid, but some of the effects observed could be
treated in terms of linear viscoelasticity, whilst others could be accounted for by nonlinear treatment where time effects were ignored.
ACKNOWLEDGMENT
It is a pleasure to acknowledge the support of the Advanced Research Projects Agency,
Department of Defense under A R P A Contract Sd-86 with Brown University, Providence,

R.I.
REFERENCES
I. International Symposium on Stress Waves in Materials x96o (N. Davids, ed.). New York:
Interscience.
2. Stress Waves in/Inelastic Solids. Proceedings of IUTAM Symposium held at Brown University
x963. Springer Verlag (in press).
3. H. KOLSKYI953 Stress Waves in Solids. Oxford : Clarendon Press (Dover reprint in press).
4. J. S. RiNmt~r and J. PEARSON X959 Behavior of Metals under Impulsive Loads. Cleveland,
Ohio : Amer. Soe. for Metals.
5. W. M. EWXNC,W. S. JARt)~rSKYand F. PRESSI957 Elastic waves in Layered Media. New York:
McGraw-Hill
6. W. GOLDSMXTItX960Impact. London: Edward Arnold.
7. R. M. DAVI~ x956 Surveys in Mechanics, G. I. Taylor 7oth Anniversary Volume. Cambridge
University Press. See p. 64.
8. K. B. BaOBERC 1956 Shock Waves in Elastic and Elastic-plastic Media. Stockholm: Report
No. Io9, Kungl. Forif. Befast. Forsk.
9. H. N. ABm~MSON,H. J. PLASSand E. A. RIPPERCEa 1958 Adv. appl. ~VIech. 5, t I I .
to. H. KOOKy x958 Appl. Mech. Rev. xx, 465.
xI. S. C. HUNTER 1959 Progress in Solid Mechanics (I. N. Sneddon and R. Hill, eds.). North
Holland Publishing Co. p. 3x2. W. A. GREEN X959 ibid. x, p. 225.
I3. J. W. O. Cm~cGs 196x ibid. ~, p. 143.
x4. M. J. P. MUSCaXvE x96x ibid. % p. 63.
x5. H. KOLSKYI960 Structural Mechanics (J. N. Goodier and N. J. Hoff, eds.). London : Pergamon
Press. See p. 233.
x6. J. MmLOWITZ X960,4ppl. Mech. Rev. 3, 865.
t7. H. KOLSKY I962 Experimental Techniques in Shock and Vibration (W. J. Worley, ed.). New
York : A.S.M.E. See p. i x.
x8. H. LAMa x89o Proc. Lond. math. Soc. 2x, 85 ; x917 Proc. roy. Soc. A93, xI4.
19. L. POeHHAMMERi876J, reine angew. Math. 8x, 324.
20. C. CmteE x889 Trans. Camb. phil. Soc. x4, 250 ; 1889 Quart. J. pure and appl. Math. 23, 335.
2x. J. GHOSH I923 Bull. Calcutta math. Soc. I3, 217; 1924 ibid. x4, 31.
22. E. GreBE and E. BLECHSCHMIDT1933 Ann. Phys., Leipzig 18, 417 .
23. G. HERMANNand I. MIRSKY 19569qt. appl. Mech. 23, 563 ; I958 ibid. 25, 9724. J. E. GmmNSPANI96oJ. aero. Space S d . 27, 3725. S. TIMOSHENKOI921 Phil. Mag. 41,744.
26. R. D. MINDLIN and G. HERMANN 1952 Proc. Ist U.S. nat. Congr. appl. Mech. New York:
A.S.M.E. See p. I87.
27. R. D. MINDLINand H. D. McNtVEN 196oj. appl. Mech. 27, 145.
28. A. E. H. Love 1928 A Treatise on the Mathematical Theory of Elastidty. Cambridge University
Press, 4th edition. See p. 29o.

I I0

29.
3 o.
31 .
32.
33.
34.
3536.
37.
38.
39.
40.
4 I.
42.
43.
444546.
4748.
4950.
5I.
5z.
53.
54.
55.
56.
57.
58.
59.
60.
6i.
62.
63.
64.
65.
66.
67.
68.
69.
7 o.
71.
72.
73.
74.
7576.
77.
78.
79.

II. KOLSKY

J. ADEM 1954 Ouart. appl. :$lath. x2, 261.


R. 1\I. DAVIES x948 Phil. Trans. ,4, 24 o, 375S. PE'rEr~.~.qON I953 Trans. Roy. Inst. Tech. Stockholm, No. 62.
E. A. Rn'PERGER I953 Proc. Ist :VIidwestern Conf. on Solid !Vlech. U r b a n a , Ill. See p. 29.
I). Y. HSIF.H and H. KOl..SKY I958 Proc. phys. Soc. 7 x, 6o8.
J. ~IIKLOWITZ and ('. R. NmEWANGER I 9 5 7 J - appl. Mech. 24, 240
C. W. Ct:R'rls I 9 5 4 J . appl. Phys. 25, 9z8.
L. Y. T u , J. W. BRF.NNA.'q and J. A. SAUER 1 9 5 5 J . acoust. Soc. Hmer. 27, 5 5 o
E. G. STANFORI) 1950 Nuovo ('ira. Suppl. No. 2, 7, 332.
l). S. I h;c;nF_q, \V. 1,. PONDROM a n d R. I,. XiIMg i949 Phys. Rc~. 75, 1552
t l . KOLSKY I954 Phil. Mag. 45, 7 I2.
I). M. CUNNINGIL.'tM a n d W. (}OI.DSMITII I 9 5 6 J . appl. :~lech. a3, 612.
E. A. RIPPERGEI~ and 11. \V. ABRAMSON I 9 5 1 J . appl. :~tech. 24, 43 I
J. D. OWEN a n d R. M. DAWrS I949 Nature (Lond.) I64, 752.
J. M. KaAH:T I 9 5 5 J . appl. Phys. 26, I248.
I'. SAUTER 1950 Z. angew. Math. Mech. 3 o, 2o 3.
J. N. GOODn:.R a n d R. E. D. BISHOP I952 J . appl. Phys. 23, 124.
F. ('. ROESLER x955 Phil. Mag. 46, 517.
11. SCHARDIN I950 Glastech. Ber. 23, 1, 67, 325.
D. G. CHRISTIE I955 Phil. Mag. 46, 527.
'I'. I). NOR'I'HWOOD a n d I). V. ANDEP~SON 1954 Bull. seismol. Soc. Amer. geophys. Un. 35, 969.
L. KNor'oFV 1954 Trans. /tmcr. geophys. Un. 3 5 , 9 6 9 .
J. OLIVER, F. PRESS a n d M. |'wINe; r954 Geophysics I9, 2o2.
J. \V. C. SHERWOOD t958 t'roc, phys. Soc. 7 x, 207.
R. l". I). BISHOP 1953 Quart. J. Mech. 6, 25o.
E. A. I~dPPERGI'R a n d H. N. ABP~.xlsox I957 Proc. 3rd Midwestern Conf. on Solid Alechs. U n i v .
of M i c h i g a n . See p. 135.
r). R. BLANI) 1960 The Theory of Linear Viscoelasticity. L o n d o n : P e r g a m o n .
l,. BOLTZMANN 1876 Pogg. Ann Eng. 7, 624.
I I. KOLSK'C I959 Progress in Non-destructive Testing 2, 29. L o n d o n : H e y w o o d .
ii. KOLSKY ~956 Phil. ;l~lag. I, 693.
K. W. IIILLIER 1949 Proc. phys. Soc. 62B, I 11.
| I. KOLSKV and S. S. LEE 196Z Brown University Tech. Rep. No. 5, Contract. N o n r 562 (30).
L. I1. DONNELL I93 o Trans. II.S.M.E. 52, ~53.
T . VON KARMAN a n d P. F. I.)t:WEZ 1 9 5 o J . appl. Phys. 2x, 987.
P. E. I)I:WEZ a n d I). S. Ct.ARK 1947 Proc. Amer. Soc. Test. Mater. 47, 5o2.
J. F. BELL 1956Johns Ilopkins Univ. tech. Rep. No. 4, A r m y C o n t r . No. D A - 3 6 - o 3 4 - O R D . , 1363.
t'. J. STEI~XGLASSand 1). A. S'rUAItT 1953 J . appl. ~,Vlech. 2o, 427.
B. E. K. ALTER and C. \V. CURTIS x 9 5 6 J , appl. Phys. 27, Io79.
L. t"..XhLVERN 1 9 5 I J . appl. ,~Iech. 18, 2o3.
II. G . HOPKINS 196I Appl. Alech. Rcv. x4, 417
l[. KOLSKV I949 Proc. phys. Soc. 6zB, 676.
E. D. U. DAVn"~ a n d S. C. HUNTER I 9 6 3 J . Mech. Phys. Solids IX, I55.
1I. KOLSKY and I,. S. I)Ot:CH I 9 6 2 J . . ~ l e c h . Phys. Solids to, ~95.
M. P. WHITE and I,. GRn:vIS ~947J. appl. 2~/lech. x 5 , 2 5 6 .
1". H. LEt.: and S. J. Tuppi.:R I 9 5 4 J . appl. Mech. 2z, 63.
K. \V. thLLn.:R a n d H. KOLSKY 1949 Proc. phys. Soc. 62B, t11.
P- \V. I]RIDC;.XlAN ~93 ! The Physics of High Pressures. L o n d o n : Bell.
I). C. PACK, \\r. ~I. EVANS a n d I | . J. JAMES 1948 Proc. phys. Soc. 6o, ~.
(;. 1". I)UVALL ~962 Appl. 3lech. Rev. I5, 849.
II. KOLSKV ~959 Fracture ( A v c r b a c h et al., eds.). T e c h n o l o g y Press a n d Wiley. See p. z8~.
S. R. BODNEa and | t. KOLSKY 1958 Proc. 3rd U.S..Nat. Congr. appl. Mech. N e w York : A . S . M . E .
See p. 495.

You might also like