You are on page 1of 317

Michael Zhuravkov

Yongtao Lyu
Eduard Starovoitov

Mechanics
of Solid
Deformable
Body
Mechanics of Solid Deformable Body
Michael Zhuravkov · Yongtao Lyu ·
Eduard Starovoitov

Mechanics of Solid
Deformable Body
Michael Zhuravkov Yongtao Lyu
Theoretical and Applied Mechanics Department of Engineering Mechanics
Belarusian State University DUT-BSU Joint Institute
Minsk, Belarus Dalian University of Technology
Dalian, Liaoning, China
Eduard Starovoitov
Construction Mechanics
Belarusian State University of Transport
Gomel, Belarus

ISBN 978-981-19-8409-9 ISBN 978-981-19-8410-5 (eBook)


https://doi.org/10.1007/978-981-19-8410-5

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Singapore Pte Ltd. 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

In this textbook, we consider the problems connecting with selecting and constructing
mechanics-mathematical models that adequately describe the stress–strain state of
deformed solids.
We do not describe detailed approaches and methods for solving specific classes
of boundary problems of deformable solids. These issues are set out in the relevant
special publications.
This publication is compiled on the basis of lecture courses for students of the joint
institutes “DTU-BSU” and “BSU-DTU” as well as textbooks and lecture courses
compiled by the authors for students of Belarussian State University, among them:
Zhuravkov M. A. Mathematical modeling of deformation processes in solid
deformable media (on the example of problems of rock mechanics). Minsk. Pub.
by Belarusian State University, 2002. 456 p. (in Russian).
Zhuravkov M. A. Fundamental solutions of the theory of elasticity and some of
their applications in geomechanics, soil and base mechanics. Course of lectures.
Minsk. Pub. by Belarusian State University, 2008. 247 p. (in Russian).
Zhuravkov M. A., Starovoitov E. I. Continuum Mechanics. Theory of Elasticity
and Plasticity. Coursebook. Minsk. Pub. by Belarusian State University, 2011.
543 p. (Classical University Edition). (in Russian).
Zhuravkov M. A., Starovoitov E. I. Mathematical models of solid mechanics.
Coursebook. Minsk. Pub. by Belarusian State University, 2021. 535 p. (Classical
University Edition). (in Russian).

Minsk, Belarus Michael Zhuravkov


Dalian, China Yongtao Lyu
Gomel, Belarus Eduard Starovoitov

v
Introduction

Deformed Solid Body Mechanics or Solid Mechanics (SM) is a collection of


disciplines that study the stress–strain state (SST) of deformed solids under the
various physical laws of material behavior.
The main “classical” sections of SM are the theories of elasticity, plasticity, and
viscoelasticity. This can also include material strength and structural mechanics.
The difference among these sections of mechanics lies in the objects under consid-
eration, assumptions and physical equations defining material behavior, and so on.
In the theories of elasticity, plasticity, and viscoelasticity, various physical laws that
establish the connection between stresses and strains are used. Often these sections
of SM are divided into mathematical and applied theories. Therefore, for example,
the mathematical theory of elasticity does not use any deformation hypotheses, and
the resulting equations are solved either by exact methods or by such approximate
methods that allow us to limitlessly increase the degree of approximation to an exact
solution. Therefore, one can consider the results obtained when solving problems by
the mathematical theory of elasticity can be used as a standard for evaluating the accu-
racy of various approximate theories and methods for solving similar problems. The
applied theory of elasticity differs from the mathematical one in that when solving
problems, in addition to the Hooke’s1 law, some additional hypotheses are used (the
hypothesis of flat sections for rods, straight normal for plates, and shells, etc.). In
solving problems of the applied theory of elasticity, along with accurate methods
for solving the corresponding equations, approximate methods are also used. Note
that there is no “clear boundary” between the applied theory of elasticity and the
resistance of materials.
The first mathematician who was engaged in systematic studies of the resistance of
solid bodies to destruction was Galileo.2 Although he considered solids to be inelastic
and did not know the law linking displacements and forces, his work indicated the
path that other researchers followed.

1Robert Hooke (1635–1703), English scientist, architect, and polymath.


2Galileo di Vincenzo Bonaiuti de’ Galilei (1564–1642), Italian astronomer, physicist, and engineer,
polymath.

vii
viii Introduction

The discovery of Hooke’s law in 1660 and the establishment of Navier’s3 general
equations in 1821 represent two important milestones in the further development of
the theory that began with Galileo. Hooke’s law provided the necessary experimental
justification for the theory. Finding common equations made it possible to reduce all
problems related to small deformations of elastic bodies to mathematical calculations.
In the period between the discovery of the Hooke’s law and the establishment of
general differential equations of the theory of elasticity, the interest of researchers
was focused on the problems of oscillations of rods and plates, as well as on the
stability of columns. These were, first of all, the fundamental work of J. Bernoulli,4
devoted to the definition of the elastic curve shape, and the work of Euler,5 which laid
the foundation for the research in the field of stability of elastic systems. Lagrange6
applied Euler’s theory to determine the most reliable shape of columns.
The mathematical theory of elasticity as a science has been developed in the
first half of the nineteenth century, mainly thanks to the work of French engineers
and scientists. Therefore, for the first time, equilibrium and oscillation equations
of elastic solids, assuming the body discrete molecular structure, were obtained by
Navier. He deduced not only differential equations, but also boundary conditions
that must be satisfied on the surface of the body. Lamé7 and Clapeyron8 developed
Navier’s theory in relation to engineering. They wrote a special and scientific work
on the internal equilibrium of solids, which solved the problem of stresses and strains
of a thick-walled pipe under axisymmetric loading (Lamé’s problem).
By the fall of 1822, Cauchy9 discovered most of the main elements of the pure
theory of elasticity. He introduced the concept of stress and strain at a point. He
showed that they (stress and strain) can be defined by six relevant components. Based
on the hypothesis of a continuous and homogeneous solid structure, Cauchy obtained
equations of motion (or equilibrium). He first introduced two elastic constants into the

3 Claude-Louis Navier (1785–1836), French mechanical engineer, a physicist whose work was
specialized in continuum mechanics. First introduced the concept of stress.
4 Jacob Bernoulli (1655–1705), one of the many prominent mathematicians in the Bernoulli family.

He was an early proponent of Leibnizian calculus and sided with Gottfried Wilhelm Leibniz during
the Leibniz–Newton calculus controversy.
5 Leonhard Euler (1707–1783), Swiss mathematician, physicist, astronomer, geographer, logician,

and engineer who made important and influential discoveries in many branches of mathematics. He
is also known for his work in mechanics, fluid dynamics, optics, and astronomy.
6 Joseph-Louis Lagrange (1736–1813), Italian mathematician and astronomer, later naturalized

French. He made significant contributions to the fields of analysis, number theory, and both classical
and celestial mechanics.
7 Gabriel Lamé (1795–1870), French mathematician who contributed to the theory of partial differ-

ential equations by the use of curvilinear coordinates and the mathematical theory of elasticity (for
which linear elasticity and finite strain theory elaborate the mathematical abstractions).
8 Benoît Paul Émile Clapeyron (1799–1864), French engineer and physicist, one of the founders of

thermodynamics.
9 Baron Augustin-Louis Cauchy (1789–1857), French mathematician, engineer, and physicist who

made pioneering contributions to several branches of mathematics, including mathematical analysis


and continuum mechanics.
Introduction ix

equations of elastic theory, while Navier’s equations contained only one. Relations
that link small strains and displacements are named after him.
A significant contribution to the development of the theory of elasticity belongs
to Saint-Venan.10 He proposed a new approach for solving a number of applications
(the semi-reverse method of Saint-Venan). Using this method, important problems
were solved regarding the bending and torsion of a bar with a non-circular cross
section. He also owns research on oscillation, shock, and plasticity theory.
In the second half of the nineteenth century, Kirchhof11 formulated the main equa-
tions of thin rod theory, which laid the foundation for the development of methods
for calculating elastic springs. In addition, he developed a consecutive theory of thin
plates. The first attempt in this direction was made by Lagrange and Sophie Germain12
in 1814, but they were not able to correctly formulate the boundary conditions.
At the end of the nineteenth century, Aron and Love13 gave the first version of
the equations of shell theory of the model, based on the application of the hypothesis
of a non-deformability of the normal rectilinear element. Bussinesk14 investigated
the problem of determining stresses in an elastic body under the influence of a
concentrated force. These studies allowed Hertz15 to formulate the problem about
the interaction of two elastic bodies in contact.
An important role in the development of the theory of elasticity was played by
the work of Russian scientists. The fundamental results of the development of the
principle of possible displacements, the theory of impact, as well as the integration
of dynamics equations belong to Ostrogradsky.16 Gadolin17 studied the stresses in
multilayer cylinders, thereby building the basis for the design of artillery barrels.
Zhuravsky18 formulated the modern theory of bending beams. Significant progress
in the construction of methods for solving 2D problems of the theory of elasticity is

10 Adhémar Jean Claude Barré de Saint-Venant (1797–1886), a mechanician and mathematician


who contributed to early stress analysis and also developed the unsteady open channel flow shallow
water equations, also known as the Saint-Venant equations.
11 Gustav Robert Kirchhoff (1824–1887), German physicist who contributed to the mechanics,

mathematical physics, electricity, and spectral analysis.


12 Marie-Sophie Germain (1776–1831), French mathematician, physicist, and philosopher. One of

the pioneers of elasticity theory.


13 Augustus Edward Hough Love (1863–1940), a mathematician famous for his work on the

mathematical theory of elasticity. He also worked on wave propagation.


14 Joseph Valentin Boussinesq (1842–1929), French mathematician and physicist who made

significant contributions to the theory of hydrodynamics, vibration, light, heat, and theory of
elasticity.
15 Heinrich Rudolf Hertz (1857–1894), a German physicist who first conclusively proved the

existence of the electromagnetic waves predicted by James Clerk Maxwell’s equations of


electromagnetism.
16 Ostrogradsky Mikhail Vasilyevich (1801–1862), Russian mathematician, mechanician, and

physicist. Performed important studies on integral calculus.


17 Gadolin Aksel Vilgelmovich (1828–1892), Russian scientist, developed the theory of fastening

the barrels of artillery guns; works on physics, metal processing.


18 Zhuravsky Dmitriy Ivanovich (1821–1891), Russian scientist and engineer, founder of a school

in the field of construction mechanics and bridge construction.


x Introduction

associated with the names of Kolosov19 and Muskhelishvili,20 who first applied the
method based on the use of complex variable functions. Bubnov21 solved a number
of problems about a plate bending. Fundamental studies on the theory of plates and
shells, vibrations of rods taking into account the influence of shear deformations
were carried out by Tymoshenko.22 Subsequently, many problems were solved by
the energy method proposed by him. Galerkin23 completed a series of studies on
the theory of bending thin plates, thick plates, and the theory of shells. To derive
the equations of shell theory, he apparently first applied the equations of the three-
dimensional theory of elasticity.
Papkovich first proposed to build a solution to the problems of the theory of
elasticity of displacements in the form of harmonic functions. He also performed
studies of general stability theorems of elastic systems. In addition, he solved a large
number of problems about bending plates under various boundary conditions.
Vlasov,24 Novozhilov25 and Rabotnov26 made a great contribution to the devel-
opment of the general theory of shells and other sections of the theory of elasticity.
Therefore, Vlasov is the founder of a new scientific discipline—the construction
mechanics of shells.
The theory of plasticity, which studies irreversible deformations of solid bodies,
as an independent section of mechanics, began “its history” about a hundred and
fifty years ago. The foundations of the theory of plastic flow whose creation was
aimed at describing metal forming processes were formulated in the first publication
of Saint-Venan (1868–1871).

19 Kolosov Guriy Vasikyevich (1867–1936), Russian, Soviet mechanic; works on mathematics,

theory of elasticity, machine theory, and mechanisms.


20 Muskhelishvili Nikolai Ivanovich (1891–1976), a renowned Soviet Georgian mathematician,

physicist, and engineer. Muskhelisvili conducted fundamental research on the theories of phys-
ical elasticity, integral equations, boundary value problems, and others.
21 Bubnov Ivan Grigorievich (1872–1919), Russian engineer, developed the basics of the ship’s

construction mechanics, the designer of two submarines.


22 Tymoshenko Stepan Prokofyevich (1878–1972), Ukrainian, Russian, and later, an American

engineer and academician. He is considered to be the father of modern engineering mechanics.


23 Galerkin Boris Grigoryevich (1871–1945), Soviet mathematician and an engineer. One of the

creators of the theory of bending of plates, his works contributed to the introduction of mathematical
methods in engineering research.
24 Vlasov Vasiliy Zaharovich (1906–1958), Russian, Soviet scientist in the field of mechanics, works

on the resistance of materials, construction mechanics, the theory of elasticity.


25 Novozhilov Valentin Valentinovich (1910–1987), Russian, Soviet mechanic; works on the theory

of elasticity, plasticity, and calculation of shells and ship structures.


26 Rabotnov Uriy Nikolaevich (1914–1985), Russian and Soviet mechanical scientist; works on the

theory of shells, the theory of plasticity, creep, and destruction of materials.


Introduction xi

At the beginning of the twentieth century, Karman,27 Huber,28 Mises,29 Nadai,30


Henki,31 and others put forward new concepts and theories that, although they did not
solve problems, expanded the range of ideas. In the 1920s and 1930s, in addition to
propose different versions of plasticity theory, important fundamental experimental
studies were carried out. It was an important stage in the development of plasticity
theory. However, by the end of the 1930s–first years of the 1940s, a situation devel-
oped when experiments could confirm one theory and at the same time refute other
theories. And the opposite, results from other experiments indicated otherwise.
Clarity was introduced by Ilyushin32 (1943–1945), which pointed out the need for
a clear distinction between the nature of deformation processes (simple and complex
deformation). As a result of these researches, Ilyushin developed the theory of small
elastoplastic deformations.
In the early 1950s, various theories of plasticity were proposed under arbitrary
complex loading. These approaches took the form of three theories: modern flow
theory, sliding theory, and the general theory of elastoplastic deformations.
The construction of the general mathematical deformation theory of plasticity
was based on the isotropy postulate formulated by Ilyushin. The basis for the further
development of the theory of the flow of elastoplastic bodies was the Drucker 33
hardening postulate on the non-negativity of the external forces work in a closed
cycle of plastic loading.
Models of viscoelastic behavior of materials began to actively develop in the
second half of the last century. In many materials under operating conditions, the
law of connection between force and displacement depends significantly on time.
This dependence is exerted in the fact that, for example, under constant loading, the
deformations do not remain constant, but increase. On the other hand, if the body is
subject to deformation and some bonds keep the deformation unchanged, the reaction
of the bonds decreases with time, “relaxed”.

27 Theodore von Kármán (1881–1963), Hungarian-American mathematician, aerospace engineer,


and physicist who was active primarily in the fields of aeronautics and astronautics.
28 Hyber Maximilian Tytus (1872–1950), scientist in the field of mechanics; works in the field of

material resistance, the theory of elasticity, the theory of plasticity, proposed a criterion for the
plasticity of the potential energy of shape change.
29 Richard Edler von Mises (1883–1953), an Austrian scientist and mathematician who worked on

solid mechanics, fluid mechanics, aerodynamics, aeronautics, statistics, and probability theory.
30 Nadai Arpad Ludwig (1883–1963), scientist mechanic, engineer; works in the field of the theory

of elasticity, plasticity, together with Lode introduced the Nadai-Lode parameter.


31 Henki Heinrich (1885–1951), scientist mechanic, engineer; the most famous works in the field

of sliding theory, plasticity, and rheology.


32 Ilyushin Alexey Antonovich (1911–1998), Russian and Soviet scientist mechanic, one of the

founders of the theories of plasticity, viscoelasticity, and other sections of mechanics.


33 Daniel Charles Drucker (1918–2001) was an American civil and mechanical engineer. Drucker

was known as an authority on the theory of plasticity in the field of applied mechanics. His key
contributions to the field of plasticity include the concept of material stability described by the
Drucker stability postulates and the Drucker–Prager yield criterion.
xii Introduction

The initial development of the modern theory of viscoelasticity is associated with


the names of L. Boltzmann,34 J. Maxwell,35 V. Kelvin, and36 W. Voigt.37 Subse-
quently, more complex models were created from elastic and viscous elements
combined in various ways.
Many achievements of the modern theory of viscoelasticity are associated with
the works of A. A. Ilyushin, A. Yu. Ishlinsky,38 Yu. N. Rabotnov, and other scientists.

34 Ludwig Eduard Boltzmann (1844–1906), Austrian physicist and philosopher. His greatest
achievements were the development of statistical mechanics and the statistical explanation of the
second law of thermodynamics.
35 James Clerk Maxwell (1831–1879), Scottish scientist in the field of mathematical physics.
36 William Thomson, first Baron Kelvin (1824–1907), British mathematical physicist, and engineer.

He did important work in the mathematical analysis of electricity and formulation of the first and
second laws of thermodynamics and did much to unify the emerging discipline of physics in its
modern form.
37 Woldemar Voigt (1850–1919), German physics. He worked on crystal physics, thermodynamics,

and electro-optics.
38 Ishlinsky Alexander Yulievich (1913–2003), Russian and Soviet mechanical scientist, one of the

founders of the theories of plasticity, viscoelasticity, and other sections of mechanics.


Contents

1 Statements of Boundary Problems of Solid Mechanics . . . . . . . . . . . . 1


2 Basic Concepts of Stress–Strain State Theory . . . . . . . . . . . . . . . . . . . . 9
2.1 Defining Concepts of Continuum Mechanics . . . . . . . . . . . . . . . . . 9
2.2 Stress Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Stress Tensor Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Equilibrium Equations of Deformed Solid . . . . . . . . . . . . . . . . . . . 16
2.5 Equilibrium Conditions at the Boundary . . . . . . . . . . . . . . . . . . . . . 18
2.6 Principal Axes and Principal Values of Stress Tensor . . . . . . . . . . 18
2.7 Maximal Tangent Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.8 Stress Deviator and Spherical Part of Stress Tensor . . . . . . . . . . . . 24
2.9 Strains and Displacements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.10 Principal Axes and Principal Values of Strain Tensor . . . . . . . . . . 32
2.11 Strain Deviator and Spherical Part of Strain Tensor . . . . . . . . . . . . 33
2.12 Strain Compatibility Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3 Material and Solid Mechanical Characteristics (Properties) . . . . . . . 51
3.1 Main Mechanical Characteristics of Materials and Solids . . . . . . . 51
3.2 Classification of Materials by the Nature of Deformation . . . . . . . 57
3.3 Complete Diagrams of Deformation and Fracture . . . . . . . . . . . . . 58
4 Construction of Mathematical Model Problems . . . . . . . . . . . . . . . . . . 63
4.1 The System of Equations to Describe the Medium
Stress–Strain State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Physical Relationships Determining the Solid Behavior . . . . . . . . 64
5 Mathematical Models of the Theory of Elasticity . . . . . . . . . . . . . . . . . 69
5.1 Basic Concepts and Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Hooke’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.3 Clapeyron’s Formula and Clapeyron’s Theorem . . . . . . . . . . . . . . . 76
5.4 Thermoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5 The Boundary Problems of Theory of Elasticity . . . . . . . . . . . . . . 79
5.6 The Theory of Elastic Boundary Problems in Displacements . . . . 80

xiii
xiv Contents

5.7 Boundary Problems of the Theory of Elasticity in Stresses . . . . . 83


5.8 Homogeneous Problem of the Theory of Elasticity
in Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.9 The Problem of the Setting of Theory of Elasticity . . . . . . . . . . . . 86
5.10 2D-Problems of the Theory of Elasticity . . . . . . . . . . . . . . . . . . . . . 89
6 Mathematical Models of Solid with Rheological Properties . . . . . . . . 101
6.1 Creep and Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6.2 Solid Mathematical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.2.1 Structural Rheological Models . . . . . . . . . . . . . . . . . . . . . . 103
6.2.2 Creep Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.2.3 General Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.3 Rheological Equations of Linear Viscoelasticity . . . . . . . . . . . . . . 128
6.3.1 General Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.3.2 Examples of Creep and Relaxation Kernels . . . . . . . . . . . 129
6.3.3 Volterra Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.4 Equations of Mechanics of Linear-Hereditary Media . . . . . . . . . . 137
7 Mathematical Models of Plasticity Theory . . . . . . . . . . . . . . . . . . . . . . . 149
7.1 Solid-Plastic Behavior Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
7.2 Material Plasticity During Tension and Compression . . . . . . . . . . 150
7.3 Elastic–Plastic Models of Material Behavior . . . . . . . . . . . . . . . . . 155
7.4 Viscoplasticity Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
7.5 Plasticity Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
7.6 Simple and Complex Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
7.7 Hypotheses of the Small Elastoplastic Deformation Theory . . . . . 164
7.8 Theory of Problems of Small Elastic–Plastic Deformations . . . . . 168
7.9 The Method of Elasticity Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.10 Plastic Flow Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
7.11 Relationship Between Plastic Flow and Theory of Plasticity . . . . 179
7.12 Formula and General Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
7.12.1 Formula of the Problems of Plasticity Theory . . . . . . . . . 180
7.12.2 General Methods of Solving the Problems
of Plasticity Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
8 Fundamental Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
8.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
8.2 Boussinesq’s Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
8.3 Hertz’s Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.4 Flamant’s Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
8.4.1 Flamant’s Problem for a Half-Space . . . . . . . . . . . . . . . . . 209
8.4.2 Flamant’s Problem for a Half-Plane . . . . . . . . . . . . . . . . . 212
8.5 Kelvin’s Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
8.6 Cerrutti’s Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
8.7 General Case of Point Load Action on Surface of Falf-Space . . . 223
Contents xv

8.8 Mindlin’s Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224


8.9 Some Generalizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
9 Dynamic Problems of Solid Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . 233
9.1 General Concepts and Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
9.2 Wave Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
9.3 Model Problems of Dynamics of Elastic Body . . . . . . . . . . . . . . . . 237
9.3.1 General Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
9.3.2 Cauchy Problem and Boundary Problems . . . . . . . . . . . . 240
9.4 Stokes Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
9.5 Natural and Forced Harmonic Oscillations . . . . . . . . . . . . . . . . . . . 246
9.5.1 Natural Oscillations of Elastic Bodies . . . . . . . . . . . . . . . . 247
9.5.2 Forced Oscillations of Elastic Bodies . . . . . . . . . . . . . . . . 249
9.6 Propagation of Shock Waves in Unlimited Elastic Bodies . . . . . . 251
9.7 Progressive Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
9.8 Rayleigh Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
9.9 Progressive Waves in a Plane Layer . . . . . . . . . . . . . . . . . . . . . . . . . 258
9.10 Semi-plane Under Action of Moving Surface Force . . . . . . . . . . . 263
9.11 Model Problems on Perturbation Propagation in Elastic
Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
9.12 Model Problems on Non-stationary Perturbation
Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
9.13 Model Problems on Volume Perturbation Propagation . . . . . . . . . 281
10 Mathematical Models of Special Classes of Solid Mechanics
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
10.1 Solving Problems for Heterogeneous Nonlinear Elastic
Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
10.1.1 Solving Problems for Physically Nonlinear Elastic
Medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
10.1.2 Solution of Problems for Orthotropic Solids . . . . . . . . . . 290
10.2 Solving the Problems of Linear Theory of Viscoelasticity
for Inhomogeneous Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
10.3 The Theory of Creep of Isotropic Hardening Media . . . . . . . . . . . 294
10.3.1 Consider the Situation of a Complex Stress State . . . . . . 299
10.4 Bilinear Theory of Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
10.5 Nonlinear Viscoelastic Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
10.6 Method of Successive Approximations in Viscoplastic
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
Chapter 1
Statements of Boundary Problems
of Solid Mechanics

Solid Mechanics (Deformed Solid Body Mechanics) (SM), like the Continuum
Mechanics (CM), is a phenomenological science. This means that it is based on an
axiomatic approach, although not as complete as mathematics.
The improvement and development of study methods for mechanical processes
and phenomena are characterized by the aiming to find a solution in the most general
form. This naturally requires a significant complication for the used mathematical
methods, which in turn leads to additional requirements for functions describing the
mechanical processes and phenomena being studied.
Today, success in solving the problem of qualitative and quantitative correspon-
dence of real mechanical processes and their mathematical description is impressive.
Therefore, mathematical and computer modeling has become the basic elements for
solving mechanics problems.
Despite the fact that the general plan for solving the boundary problems of SM
is quite clear, its implementation is accompanied by great difficulties. In general, it
is impossible to solve the equations describing the state and behavior of deformable
solids under various loads. Analytical solutions exist only for a limited number
of applications. However, even these solutions are of great value, since there is a
standard with which you can compare approximate solutions obtained as a result of
the introduction of certain simplified hypotheses.
Therefore, in mechanics, we’re dealing with modeling. In particular, in SM, we
are dealing with modeling the deformation processes of solid bodies. SM operates
models with elastic body, plastic body, etc., although real medium is described using
these models only under certain assumptions. A continuous medium is introduced
to describe discrete physical objects so that a known effective mathematical analysis
apparatus can be used. Just as any set becomes a space only after introducing some
structures, continuous medium needs to be introduced by axioms and postulates to
become the object of constructing models.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 1
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_1
2 1 Statements of Boundary Problems of Solid Mechanics

As B.E. Pobedrya1 noted, the continuity principle used to introduce a continuum


“does not solve all problems”. After setting the continuum mechanics problem, its
solution is carried out using computational mathematics. To do this, the problem must
be “transformed” into an algebraic problem, that is, discretization. Furthermore, in
order to analyze the solution obtained and compare it with the experimental results,
it is necessary to continue the problem again.
Thus, the processes of discretization and continuity of CM problems and, in
particular, SM problems are repeated several times at different levels.
When we consider mathematical models for real complex mechanical processes, it
is extremely difficult to achieve the high quantitative accuracy. One of the reasons for
this situation is, for example, the discrepancy between the mechanical and strength
characteristics used for calculations and the real characteristics of objects. The next
reason is that it is impossible to construct a mechanical-mathematical model that
accurately reflects all the features of the real process, etc.
The main problem of the modeling is that, based on the results of model studies,
it is possible to give the necessary answers about the qualitative and quantitative
features and effects of the studied phenomenon in real conditions.
Modeling in the general case can be divided into three types: physical, mathemat-
ical, and functional modeling.
This publication relates to the mathematical modeling of the stress–strain
state of solids under conditions of various forces and kinematic loadings. Math-
ematical modeling is used to study those processes that can be described mathe-
matically (i.e., for which mathematical models can be built), in such a way that the
mathematical problem of constructed model is solvable.
Currently, due to the active introduction of computer tools in all areas of human
research and production, the approaches and ideologies of scientific research and
engineering work are radically changing. The computer today serves as an intelligent
assistant. Now experts, in the majority, don’t put their attentions on the development
of approaches to the solution of problems, but deal with a problem of adequacy of
a model problem and real process. At the same time, the methods for solving the
model problems can be very “bulky and heavy”. Since today’s tasks are solved using
specialized software, this circumstance is not a determining factor.
As a rule, solving a large number of applied mechanical problems requires passive
or active experiments.
The main disadvantage of a passive experiment (by this term, we mean studies or
observations “in-situ”) is the impossibility of sufficient variation of the input param-
eters, which limits the use of the results obtained to the framework of the specific
conditions in which the studies were performed. To eliminate this drawback, the real
process is replaced by a model, and subsequent studies are carried out with a variation
of input parameters with the help of this model (we will call it active experiment).
The most widely used technologies of an active experiment are approaches based on
the use of physical and mathematical models.

1Pobedrya Boris Efimovich (1937–2016), Soviet and Russian scientist in the field of mechanics;
works in the field of the theory of elasticity, viscoelasticity, plasticity, numerical methods.
1 Statements of Boundary Problems of Solid Mechanics 3

Before the active introduction of computer technologies, mathematical modeling


methods were not widely used in the study of complex applied processes due to
the laboriousness and unscheduled impossibility of conducting real calculations in
accordance with the constructed mathematical model. In this regard, physical models
were more preferred and common. However, physical models, in turn, in addition to
the positive aspects, have a large number of negative factors that prevent the spread
of physical modeling as a universal technology.
Therefore, it is necessary to clearly represent the areas of approaches which
are effectively used and limited to real-world process modeling.
Today, there is a procedure for actively replacing physical modeling with mathe-
matically based computer technologies in accordance with predetermined physical
equations describing the behavior of the medium.
Thus, today one of the most important problems is the development and adap-
tation of modern advanced approaches and methods of mathematical modeling to
perform computer modeling for a wide range of applied mechanical processes. Real
mechanical processes are very diverse and complex in terms of adequately qualitative
and quantitative descriptions based on the mathematical modeling.
Two approaches to the theoretical construction in fundamental and applied
sciences are distinguished—phenomenology and structuralism.
Phenomenological models are based on empirical data of object behavior. The
structural approach consists of the development of models that allow you to describe
and explain the phenomena based on the internal structure of the objects. Note
that these two approaches are closely interrelated. Therefore, in addition to these
two approaches, the approach called structural-phenomenological approach has
become widespread. The essence of this approach is that the standard phenomenolog-
ical equations and criteria in deformed solid mechanics are considered at two levels:
microscopic and macroscopic, displaying behavior of non-homogeneous materials
as homogeneous materials with effective properties.
Theoretical modeling concerns two main points in the SM:
a) Constructing models describing the physical state of the material of which the
solid consists;
b) Setting boundary problems taking into account constitutive (defining) equations
of material behavior.
The mathematical model of the original object is a system of equations in the
general sense of this term. The same mathematical model can describe different
physical processes (models). However, when studying the same real objects, funda-
mentally different models can be used. Therefore, the choice of model is very signif-
icant in the researches. The construction of different models of the same object is
aimed at describing a different detailing of the processes or phenomena.
Relative to the selected system of characteristics, the adequacy of the math-
ematical model to the studying object (process, phenomenon, etc.) belongs to
the basic requirements for the mathematical model. At the same time, adequacy
should be considered on two sides: the correct qualitative description of the object
by the selected characteristics and the correct quantitative description of the object
4 1 Statements of Boundary Problems of Solid Mechanics

by the selected characteristics with a certain degree of accuracy. The adequacy of


the modeling is determined: by the study object itself, the object model, external
influences, and the selected class of responses to them, as well as the accepted level
of accuracy.
It should be remembered that universal adequacy does not exist. The adequacy of
the model should be considered only by certain features and characteristics accepted
as basis in this study. When considering the model of an object, it is necessary to find
out whether the accepted simplifications only lead to a permissible loss of accuracy
or a qualitative difference between the model and the real object. At the same time,
the legality of the application of the introduced hypotheses is also checked.
Selecting factors that you can ignore when you build up a mathematical model is
an important task. It is obvious that more complex models are required to describe
the object more adequately, since they can take into account a larger number of
factors that affect the study of characteristics. However, complicating the model, in
turn, requires complicating the mathematical apparatus. It may also happen that the
constructed systems of equations for a solution will be unjustified, in terms of conclu-
sions, time, materials, and human expenditures. In addition, such complex systems
are generally solved by approximate methods with a certain degree of accuracy, and
adequacy can be significantly lost with such a transition.
Thus, the requirement of adequacy must be considered in conjunction with the
requirements of sufficient simplicity and optimality of the model with respect to
the selected system of characteristics. The model shall be selected in such a way that
quantitative or qualitative analysis of the selected characteristics and properties can
be carried out optimally in terms of time and labor costs, and with varying accuracy
using modern research tools.
The relationship between the adequacy and simple properties of the models is
shown in Fig. 1.1, where these properties are estimated by numbers “a” (adequacy)
and “s” (simplicity), taking values from zero to one.
Obviously, only points within the hatched curved triangle are of practical interest.
It must be remembered that, in most cases, the way to choose a less adequate (but
still satisfactory) physical model was more effective and its mathematical model did
not require additional simplification.
For a large class of mechanical tasks, infinity problems are important.
On the one hand, mechanical processes under the influence of external loading
are realized only in some limited areas. Therefore, when constructing a calculation
scheme, only a part of the body is usually considered. At the same time, external
boundaries are chosen in such a way that the mechanical processes are practically
beyond their limits.
On the other hand, the real objects are almost always finite, so an infinite math-
ematical object (infinite dimensions, infinite interval, etc.) appears as a result of
simplifying the mathematical model. The real number of elements or the real size
of the interval, allowing the transition to mathematical infinity, in different problems
are very different. This choice depends on many factors and primarily on the “degree
of influence” of far elements and on the permissible error in solving the problem.
Therefore, in application problems, any infinity is countable, and when modifying
1 Statements of Boundary Problems of Solid Mechanics 5

Fig. 1.1 Qualitative diagram of the relationship of the adequacy and simplicity properties of the
model

a problem, an infinite set can become finite and vice versa. Infinity can also result
when each element loses its personality. In this case, discretization is replaced by
continuity.
The concept of “infinitesimal” is also important in the construction of mathemat-
ical models of many mechanical processes. It should be noted that there are no well-
founded general considerations and recommendations on the choice of the concepts
of “infinitely large” and “infinitely small” and the related concepts of “significantly
more”, “significantly less” still are not.
Excessively large and excessively small numbers are also of auxiliary importance.
An assessment of the accuracy of the solutions is associated with these concepts. For
example, it is obvious that it makes no sense to spend effort to determine in the rock
massif the area of disturbance with the additional stress state with an accuracy of
1 N. However, at the same time, the difference of 1 mm in displacements is important
to estimate the plate stability.
An important conclusion from the facts presented is: the choice of main char-
acteristics and their values plays a very important role in the construction of the
model. The main characteristics are the spatial and time variables that we use
when constructing a model for an object of research. After defining the main char-
acteristics (scales), all variables relative to them can be classified as variables with
“normal”, “slow” and “rapid” rate of change. Slowly changing values can be consid-
ered parametrically (for example, taken as constant), and rapidly changing values
are taken into account by averaging and introducing effective values.
Therefore, for example, a variable value is the thickness of the layerlin a layered
composite. The thickness of the layer has a characteristic range of change l∗ . The
6 1 Statements of Boundary Problems of Solid Mechanics

composite has a characteristic size d∗ . Then, as one of the main characteristics, it is


advisable to take a value equal to the ratio l∗ to d∗ : δl∗ = l∗ /d∗ . In this case,
when you build a model, all values are classified as follows: main if their respective
parameters (defined similarly δl∗ ) are comparable to δl∗ ; slow if the values of these
parameters are much smaller δl∗ ; fast if parameter values are much larger δl∗ .
If you want to build more accurately than the original model, you enter clarifying
values.
The forming process of the main variables can be represented by the diagram
shown in Fig. 1.2.
Similar reasoning concerns the duration of the effects of disturbances on the
mechanical system.
Classification of disturbances when taking into account the time of external
influence can be performed as follows:
• Disturbances change throughout the entire time interval of the mechanical process
study;
• Influence of disturbance is felt only in the integral characteristic (e.g., dynamic
impact, explosion);
• Perturbation occurred so long ago that due to the dissipation of energy, its effect
in the time interval under consideration is no longer felt.
The mathematical modeling process of mechanical processes and phenomena can
be structured as follows:
• The ultimate aim of the problem is formulated;
• The type of problem is set from initial geometric conditions;
• Physical model of the object with indication of acting kinematic and force loads
is accepted;
• Quantitative estimates of mechanical properties of materials and structural
features of the objects are determined;
• Mathematical formula of the problem is performed, the calculation scheme is built
with a setting of initial and boundary conditions;
• The method of solving the formulated mathematical problem is chosen;
• Direct solution of the model problem is performed;
• Analysis and interpretation of the obtained results of the solution are carried out.
It should be noted that if the formulas of the main aim of the study is usually not
very difficult, since it is dictated by the requests of practice, then the determination
of the necessary initial data sometimes requires special studies.
Control Questions
1. Mechanics of deformed solids is a phenomenological science. What does it mean?
2. What are the “principles of continuity and discretization” in SM?
3. What are “active and passive experiments” in SM?
4. How do you understand the terms “theoretical models of the mechanical behavior
of material” and “setting a model boundary problem taking into account the model
of material behavior”?
1 Statements of Boundary Problems of Solid Mechanics 7

Defining parameters

Defining variables Defining constants

Main parts of
defining variables

“Slow” changing “Fast” changing


variables variables

Constant values or Variables with a Averaged values of


values depending “normal” (basic) “fast” changing
on t or x rate of change variables
parametrically

Main variables

a) Corrections b) Corrections
that take into that take into c) Small parts of d) Additional
account changes account rapid defining variables Defining
in slow variables changes in parameters
over time or variables relative
space to their averages

Clarifying variables

Fig. 1.2 Main variable generation process


8 1 Statements of Boundary Problems of Solid Mechanics

5. How do you understand the requirement of the adequacy of a mathematical model


to a real object of research?
6. What are the concepts of “infinitely small” and “infinitely large” in SM?
7. What are the “basic characteristics” of concepts; variables with a “normal”,
“slow” and “rapid” rate of change; “clarifying values” in SM?
8. What are the process steps of mathematical modeling of mechanical processes
and phenomena?
Chapter 2
Basic Concepts of Stress–Strain State
Theory

2.1 Defining Concepts of Continuum Mechanics

The mechanical behavior of a solid can be very diverse and complex. In this course,
the description of the deformed solid state and behavior is based on the theory
of a continuous medium. We will ignore (but not always) the discrete structure
of the material. Therefore, it is assumed that the volume occupied by the body is
continuously filled with matter.
We consider the infinitesimal volume of material as a “particle” of a continuous
medium. At the same time, we accept that all individual particles of a continuous
medium are absolutely the same and indistinguishable. This assumption is one of
the basic concepts of the continuum mechanics.
The choice of the continuum model is based on the concepts of elementary
volume, the criteria of quasi-continuum, and quasi-heterogeneity.
The choice of elementary volume size is more influenced by the structural defects
of medium. Classification of volume as elementary volume depends on the size of the
object. The question of the legality of using continuum mechanics methods should
be solved for the specific volume of each body, taking into account its structural and
mechanical features.
The criterion of quasi-continuum can be formulated as follows:

ΔA < ε by Δa < l0 ,

where ΔA is the difference among the values of stresses, strains, or displacements


in neighboring points of the medium with an increment of coordinates; Δa, l0 are
the characteristic linear sizes of the elementary volume; ε is the permissible error in
the determination.
The criterion for considering the medium as quasi-homogeneous materials can
be taken to exceed the impact area over the size of the inhomogeneous medium by
at least a value order in linear dimensions:

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 9
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_2
10 2 Basic Concepts of Stress–Strain State Theory

Vg / Vn = 103 ,

where Vg and Vn are the volume of the impact area and heterogeneity, respectively.
Deformable solids change sizes and shapes (deformation) under the influence of
external forces (loads). Internal forces arise in deformed solids. The magnitude and
distribution of internal forces depend on the load and the body’s geometric shape.
External loads can be divided into volume, surface, and concentrated loads. The
concentrated loads can be considered as a limit case of applying surface loads on a
small part of the body surface.

2.2 Stress Tensor

We divide the body into two parts by an imaginary section: V 1 and V 2 , (Fig. 2.1a).
The surface element of section ΔS with the focal point A has a unit normal vector ν
directed to V 1 . The force vector ΔP and the moment vector ΔM represent the action
of the part V 1 body at point A on part V 2 . In the limit, when the elementary section
aims for point A, ΔS → 0, (with a fixed direction ν), the following physically
reasonable assumptions can be made:

ΔP dP ΔM
lim = = σν , lim =0 (2.1)
ΔS→0 ΔS dS ΔS→0 ΔS

The vector σν (2.1) is called the stress vector at point A. It is important that vector
σν acts on the surface element ΔS with the normal direction ν and changes when
the direction of this normal to ΔS changes.
For stress vectors acting at the same point of the section, but directed in opposite
directions, the following equality is true:

σv = −σ−v (2.2)

Fig. 2.1 To the definition of “stress on the site”


2.2 Stress Tensor 11

Fig. 2.2 Normal and


tangent components of the
stress vector

Equality (2.2) describes the effect of the body part V 2 on the part V 1 and vice
versa, (Fig. 2.1b).
Remark. The relation (2.2) can be interpreted as a direct expression of the Newton1
third law (the principle of equality of action and opposition).
The set of stress vectors σν (A) for all directions ν determines the stress state at
point A.
In the general case, the stress vector σν (A) is not directed along the normal ν. Its
projection to an arbitrary direction (defined by a unit vector) is called the component
of the stress vector in this direction. If we decompose the stress vector σν into normal
and tangent to the site dS, then we get the normal stress σn and the tangent stress
τn , (Fig. 2.2).
For modulus of these values, the ratio is true:

σ2ν = σn2 + τn2 .

It should be noted that the stresses σn and τn are not components of the vector in
the usual sense.
We will use Cartesian coordinates x 1 , x 2 , x 3 , which we will denote with the
symbol x i , while remembering that i and other Latin indices take values of 1, 2, 3.
Vectors ei denote the basis vectors of the coordinate system.
To describe the stress state at a point of solid, an elementary parallelepiped is cut
around this point, the edges of which are parallel to the coordinate axes and have
small lengths dx i , (Fig. 2.3a). Due to the small parallelepiped, it can be considered
that the stress vectors σi on its edges coincide with the stress vectors on the parallel
coordinate sites drawn through the point under consideration.
The important assumption that deformed and undeformed solid elements are
identical should be noted. This follows from the hypothesis of small strains.
The stress vectors σi acting on the parallelepiped faces can be decomposed into
normal σii (perpendicular to the faces) and tangent σij (i /= j) components (lying in
the plane of the faces), (Fig. 2.3b).

1 Newton Isaac (1643–1727), English mathematician, mechanician; founder of classical mechanics.


12 2 Basic Concepts of Stress–Strain State Theory

Fig. 2.3 Decomposition of


the stress vector in the
general case

We accept that the first subscript in the symbols indicates the normal of the section
on which this stress acts, and the second index indicates the axis on which it is parallel.
Tensile normal stresses are positive and compressive stresses are negative.
We emphasize that the values σij are not components of the vector in the usual
sense. They are measured most often in pascals (1 Pa = 1 N/m2 ).
For example, the decomposition of the stress vector σ3 (acting on the face x 3 =
constant) along the coordinate axes x i has the form:

Σ
3
σ3 = σ31 e1 + σ32 e2 + σ33 e3 = σ3k ek ≡ σ3k ek .
k=1

We accept, as in the classical course of the continuum mechanics, that when a


Latin index is found twice in a single member, a summation of this index from 1 to 3
occurs. We will not write the summation mark itself (this index is called mute). This
index can be replaced by any other indices. For example, xi yi = xk yk .
If the Greek index is repeated, then it is not summed. The index be found once
in the term is called free. With the correct spelling of any formula, each of its terms
must have the same free indices.
In the general case, for the stress vector σi , the following equality is true:
2.2 Stress Tensor 13

σi = σi j e j , (2.3)

where ej is the unit vector of coordinate axis x j .


We will actively use the concept of a metric tensor. Therefore, recall that the
metric tensor δij is defined as the scalar product of the basis vectors. In Cartesian
coordinate system ei · e j = 0, if i /= j, so:

1, i = j,
δi j = ei · e j , δi j = (2.4)
0, i /= j.

The matrix of this tensor is diagonal. Generally, the components of the metric
tensor (2.4) are called Kronecker 2 symbols.
Scalar multiplication of the ratio (2.3) by the basis vector ek gives:

σi · ek = σi j e j · ek = σi j δi j .

It follows from (2.4) that δkj = 1 only with j = k, therefore, we get:

σik = σi · ek . (2.5)

The nine stress components σij represent, in aggregate, a physical quantity


called the Cauchy stress tensor3 (T s ). This is a second-rank tensor (by the number
of indices). Write it as a matrix, like this:
⎛ ⎞
σ11 σ12 σ13
T s = (σi j ) = ⎝ σ21 σ22 σ23 ⎠.
σ31 σ32 σ33

Matrix rows contain stress tensor components parallel to one axis, and columns
contain tensor components acting on the same section.
Components σi j mean stresses at point N acting in the direction of axis i on a
section with an external normal towards axis j.
In this case, the stress vector σ→i at point N in the direction of axis i is defined as:
σ→i = σ→i {σi1 , σi2 , σi3 }.
Recall that the tensor is an invariant object, that is, it does not change when
transitioning from one coordinate system to another. Only its components change
according to a certain “tensor” law with such a transition. However, in the future, we
will sometimes call “tensor” the combination of its components.
Different symbols can be used for stress tensor components. Therefore, for
example:

2 Leopold Kronecker (1823–1891), a German mathematician who worked on number theory,


algebra, and logic.
3 The concept of “tensor” was formed from the Latin “tension”—stress.
14 2 Basic Concepts of Stress–Strain State Theory
⎛ ⎞
σx x σx y σx z
Ts = ⎝ σ yx σ yy σ yz ⎠,
σzx σzy σzz

where x, y, z are the coordinate axes.


The representation of the Karman has the form:
⎛ ⎞
σx x τx y τx z
Ts = ⎝ τ yx σ yy τ yz ⎠.
τzx τzy σzz

Thus, the stress state in the body is determined by the field of stress tensors, that
is, by setting at each point of the body nine functions σi j (x, y, z, t) (t is time), which
are related to the external normal to the site.
The stress tensor is symmetrical, i.e., σi j = σ ji .
A very important statement is the law of reciprocity: if there are two intersecting
elementary sites at the body point, then the projection of stress on the first site to
the normal to the second equals to the projection of stress on the second site to the
normal to the first (this is a more general statement than the symmetry of the stress
tensor).

2.3 Stress Tensor Properties

The components of the stress vector σν on an arbitrary inclined site, are carried
out through the observation point, and can be expressed through the stresses on the
coordinate sites (faces of the elementary parallelepiped) at this point, (Fig. 2.4a).

Fig. 2.4 To define stress on an arbitrary site with normal ν


2.3 Stress Tensor Properties 15

To prove this statement, we consider an elementary pyramid formed by coordinate


faces and an inclined site with the normal ν, (Fig. 2.4b), where σνi are projections of
a stress vector σν to coordinate axes; S is the area of the inclined site; S i are areas
of corresponding coordinate faces (subscript indicates normal to the site), which are
related to S by the following relations:

Si = Sli , (i = 1, 2, 3), (2.6)

where l i = cos(ν, x i ) are the directing cosines of the normal ν to the inclined site.
We compose equilibrium equations of forces on the faces of the selected pyramid,
which are the products of stresses on the faces on the area of the corresponding faces,
(Fig. 2.4b):

σν1 S − σ11 S1 − σ21 S2 − σ31 S3 = 0,


σν2 S − σ12 S1 − σ22 S2 − σ32 S3 = 0,
σν3 S − σ13 S1 − σ23 S2 − σ33 S3 = 0.

This system of equations takes the form in tensor symbols:

σνi S − σ ji S j = 0. (2.7)

Recall that the summation is made according to the repeated Latin index j, that
is, in the case under the consideration:σ ji S j = σ1i S1 + σ2i S2 + σ3i S3 .
We substitute the relationship (2.6) into Eq. (2.7). Reducing then by S, we obtain
formulas expressing the components of the stress vector on an arbitrary inclined site
through coordinate stresses:

σνi = σi j l j . (2.8)

Thus, the stress tensor components at the coordinate sites can fully describe the
stress state at the point.
The stress vector σν on an arbitrary inclined site drawn through a given point can
be represented as a decomposition according to the basis vectors:

σν = σνi ei ≡ σν1 e1 + σν2 e2 + σν3 e3 . (2.9)

The vector modulus σν can be expressed through the sum of the squares of its
projections using (2.9):

σν2 = σν1
2
+ σν2
2
+ σν3
2
.

The appropriate normal σn and tangent τn stresses on the same site are determined
as follows, (Fig. 2.2):
16 2 Basic Concepts of Stress–Strain State Theory

σn = σi j li l j = σ11l12 + σ22 l22 + σ33l32 + 2σ12 l1l2 + 2σ23l2 l3 + 2σ31 l3l1

τ2n = σν2 − σn2 .

2.4 Equilibrium Equations of Deformed Solid

If the deformable body is in equilibrium, then any of its part should also be in
equilibrium.
Let the body be loaded with given surface forces Rν and homogeneous volume
forces ρF (ρ is the density of the material):

ρF = ρFi ei ≡ ρF1 e1 + ρF2 e2 + ρF3 e3 .

We assume that the components of stresses σij , as well as their first partial deriva-
tives, are continuous coordinate functions. It is important that the strains are consid-
ered small, so equilibrium equations can be formulated for an undeformed body or
parts of it.
For the selected element with small edges dx i , loaded according to Fig. 2.5, the
stress components change accordingly to dσij , at the increment of coordinates by
infinitesimal value dxi change accordingly to dx i .
For example, a change of the normal stresses σ11 acting along the axis x 1 by
incrementing only dx 1 will be as follows:

∂σ11
dσ11 = σ11 (x1 + dx1 , x2 , x3 ) − σ11 (x1 , x2 , x3 ) = dx1 ≡ σ11 ,1 dx1
∂ x1

The stress σ21 acting along the axis x 1 on the other face is incremented by changing
the coordinate x 2 by the value dx 2 :

Fig. 2.5 To construct


equilibrium equations
2.4 Equilibrium Equations of Deformed Solid 17

∂σ21
dσ21 = σ11 (x1 , x2 + dx2 , x3 ) − σ11 (x1 , x2 , x3 ) = dx2 ≡ σ21 ,2 dx2
∂ x2

The stress σ31 is similarly changed: dσ31 = σ31 ,3 dx 3 .


Remark. For brevity, the partial differential operation is indicated by a comma before
the corresponding lower index:


(...) ≡ (...),i .
∂ xi

The remaining stress increments along the coordinate axes are as follows: dσαβ =
σαβ ,α dxα , (α, β = 1, 2, 3; no summation by repeating Greek index).
We obtain the known law of tangent stress parity σij = σji (i /= j)4 from the condi-
tion that the sum of moments of forces acting on the element relative to coordinate
axes is zero.
This law shows that the stress tensor has only six independent components and
its matrix is symmetrical with respect to the main diagonal.
Consider the equilibrium of the forces acting on the selected element in the direc-
tion of the x 1 axis. We project all the active forces, including volume ones, on this
axis. For this, stresses parallel to the x 1 axis are multiplied by the area of the faces on
which they act, and the force ρF 1 is multiplied by the volume of the parallelepiped
and then summed up:
( ) ( )
σ11 + σ11,1 dx1 − σ11 dx2 dx3 + σ21 + σ21,2 dx2 − σ22 dx3 dx1 +
+(σ31 + σ31 ,3 d x3 − σ31 )d x1 d x2 + ρ F1 d x1 d x2 d x3 = 0,

from which:

σ11 ,1 +σ21 ,2 +σ31 ,3 +ρF1 = 0

We take into account the symmetry of the stress tensor (σij = σji ) and obtain:

σ11 ,1 +σ12 ,2 +σ13 ,3 +ρF1 = 0

The other two directions are similar:

σ21 ,1 +σ22 ,2 +σ23 ,3 +ρF2 = 0, σ31 ,1 +σ32 ,2 +σ33 ,3 +ρF3 = 0

In general, these three equations can be written in the following tensor form:

σi j , j +ρFi = 0 (2.10)

4 Cauchy first proved in 1822 when obtaining equilibrium equations.


18 2 Basic Concepts of Stress–Strain State Theory

Equation (2.10) is the equilibrium equation of deformed solids or static


equilibrium equations.
It should be noted that in the theory of generalized media, when moment stresses
appear (for example, in the theory of the Cosser continuum5 ), the stress tensor is no
longer symmetric.

2.5 Equilibrium Conditions at the Boundary

Equilibrium Eq. (2.10) is valid everywhere within a deformable solid body. At the
boundary, that is, on the surface of the body, the equilibrium conditions in stresses
must be satisfied. This means that there must be a continuous transition of the stress
tensor to the surface load at the boundary. From (2.2) follows:

Rν = −σ−ν = σν ,

where Rν is the stress vector on the boundary specified on the surface side with the
normal ν.
We multiply this expression scalarly by ei and use the formulas (2.8). As a result,
we obtain boundary conditions in stresses:

Rνi = σi j l j (2.11)

Therefore, stresses at the boundary (surface loads) are in equilibrium with stresses
inside the body.
Equality (2.11) in coordinate form has the following form:

Rν1 = σ11 cos(ν, x1 ) + σ12 cos(ν, x2 ) + σ13 cos(ν, x3 ),


Rν2 = σ21 cos(ν, x1 ) + σ22 cos(ν, x2 ) + σ23 cos(ν, x3 ),
Rν3 = σ31 cos(ν, x1 ) + σ32 cos(ν, x2 ) + σ33 cos(ν, x3 ).

Note that the components of the normal vector are guide cosines. In the future,
we will not write the sign “ν” in the lower index.

2.6 Principal Axes and Principal Values of Stress Tensor

We take the elementary parallelepiped of a solid in the Cartesian coordinate system


and begin to rotate it around a point. The values of the stress tensor components
on its faces will change. It has been proved that there is at least one parallelepiped

5 The Cosser brothers first introduced the concept of a generalized media and studied it.
2.6 Principal Axes and Principal Values of Stress Tensor 19

Fig. 2.6 To define principal axes and principal values of stress tensor

position at which tangent stresses on its faces turn to zero, and normal stresses become
extreme, (Fig. 2.6). Coordinate axes at a given point in this position are called the
principal axes of the stress tensor. Stress components in these axes denote σ1 , σ2 , σ3
and are called the principal values of the stress tensor.
The main axes are numbered so that in the algebraic sense the following conditions
are fulfilled for the principal values: σ1 ≥ σ2 ≥ σ3 .
Matrix of the stress tensor in principal axes takes diagonal form. The components
of the stress vector on the inclined site (2.8) in the principal axes are as follows
(without summation by i):

σνi = σi li . (2.12)

Consider the procedure for determining the principal stresses values σ1 , σ2 , σ3


from the known values of the six-coordinate components of the stress tensor σij .
We use representations (2.5) and Fig. 2.4. Let the inclined site be the principal
one. Then the vector of total stress σν is directed normal ν to this site, since there are
no tangent stresses, (Fig. 2.7).
We denote the algebraic value of this principal stress as σ. Then the projections
of vector σν on the coordinate axes will be as follows: σνi = σli .
We substitute these expressions into relations (2.8) and transfer all the terms to
the left. As a result, we get such a system of equations:

Fig. 2.7 Direction of stress


vector at principal site
20 2 Basic Concepts of Stress–Strain State Theory

σi j l j − σli = 0. (2.13)

We represent the term σli in Eq. (2.13) in the form:

σli ≡ σδi j l j

Remark. This operation (li ≡ δij l j ) is called “juggling” by indices using a metric
tensor. We check this expression, for example, with i = 1:

σl1 = σδ1 j l j = σδ11l1 + σδ12 l2 + σδ13l3 = σδ11l1 = σl1 ,

where δ12 = δ13 = 0, δ11 = 1, due to (2.4).


We now substitute these relations into Eq. (2.13). As a result, we obtain a system
of three homogeneous linear algebraic relations relative to unknown cosines of the
normal, which determine the orientation of the principal site in the original coordinate
system:
( )
σi j − σ δi j l j = 0 (i, j = 1, 2, 3), (2.14)

If the solution of the received system is unique, it will be zero: li = 0, (i = 1, 2,


3). But a zero solution does not make physical sense.
For the existence of a non-zero solution, it is necessary and sufficient that the
determinant of the system is zero:
| |
| σ13 ||
| | | σ11 − σ σ12
|σi j − σδi j | ≡ | σ21 σ22 − σ σ23 | = 0 (2.15)
| |
| σ σ32 σ33 − σ |
31

This is achieved by the proper (correct) selection of the value σ. If the condition
(2.15) is satisfied, then at least one of the three equations of the system (2.14) is not
independent and is linearly expressed through the other two. Add this condition to
independent relations:

li li = 1, or l12 + l22 + l32 = 1

As a result, we get a new system of three independent equations. This system


is sufficient to find cosines of the normal of the principal site, on which the stress
equals to σ.
The equation for determining the principal stresses values is the determiner (2.15).
We disclose this equation and arrange the terms according to the power σ. As a result,
we obtain a cubic equation, which is called the age-old equation:

σ3 − σ2 J1 + σ J2 − J3 = 0,
2.7 Maximal Tangent Stresses 21

where:

J1 = σii ≡ σ11 + σ22 + σ33 = σ1 + σ2 + σ3 ,


J2 = (σii σ j j − σi j σi j )/2
= σ11 σ22 + σ22 σ33 + σ33 σ11 − σ12
2
− σ23
2
− σ31
2

= σ1 σ2 + σ2 σ3 + σ3 σ1 ,
| |
| |
|| || | σ11 σ12 σ13 |
J3 = σi j ≡ | σ21 σ22 σ23 || = σ1 σ2 σ3
|| || | (2.16)
|σ σ σ |
31 32 33

You can see that all the three roots of an age-old equation are real. They are the
principal values of the stress tensor and are indicated by σ1 , σ2 , σ3 . Their values
are determined by the nature of the external load and do not depend on the initial
orientation of the coordinate system. That is, the principal stresses are invariant with
respect to the transformation of the coordinate system. Therefore, the values of the
coefficients J 1 , J 2 , J 3 in the age-old Eq. (2.16), must remain unchanged when the
axes are rotated. In this regard, they are called the stress tensor invariants.
Planes perpendicular to the principal axes of the stress tensor is called the
principal planes.
We will follow the generally accepted rules and consider stresses positive if they
are tensile. In addition, we will number the principal stresses in descending order of
their values, that is: σ1 ≥ σ2 ≥ σ3 .
In the principal axis coordinate system, the normal stress vector σ→n =
n , x) + σ→ y cos(→
σ→x cos(→ n , y) + σ→z cos(→
n , z) is collinear with the external normal
vector n→ to the site. If the direction is the principal one, the tangent stress σ→nτ is zero
on the site perpendicular to it.
Stress tensor in principal axes has the form:
⎛ ⎞
[ ] σ1 0 0
σi j = ⎝ 0 σ2 0 ⎠, (2.17)
0 0 σ3

The stress state is called three-dimensional (triaxial) stress state, if all principal
stresses are not zero. The stress state is called two-dimensional (plane) stress state, if
one of the principal stresses is zero. The one-dimensional (linear) stress state occurs
if two principal stresses equal to zero at the same time.

2.7 Maximal Tangent Stresses

Sites are called octahedron, if they are equally inclined to the principal axes of the
stress tensor.
22 2 Basic Concepts of Stress–Strain State Theory

Fig. 2.8 Arbitrary site with


stresses in main axes

Let’s build expressions for stresses on octahedral sites. However, first, we consider
an arbitrary site in the principal axes, (Fig. 2.8). We denote the stress vector on this
site as σν , which was decomposed into the normal σn and the tangent τn components:

σν2 = σn2 + τ2n (2.18)

The projections of vector σν on the principal axes, in accordance with formulas


(2.12), take the form:

σvα = σα lα ,

where l α are guiding cosines of the normal to the site.


Then the modulus of the total stress vector on this site can be obtained from the
sum of the squares of the projections, and the normal stress value σn is obtained as
the sum of the projections of total stress components on the normal:

σv2 = σv1
2
+ σv2
2
+ σv3
2
= σi2 li2 ,
(2.19)
σn = σv li = σ1l12 + σ2 l22 + σ3l32 = σi li2

Substituting the resulting expressions for σν 2 and σn into Eq. (2.18) and consid-
ering that the sum of the cosine squares li l i = 1, we obtain the following expression
for tangent stresses on an arbitrary site:

τ2n = (σ1 − σ2 )2 l12 l22 + (σ2 − σ3 )2 l22 l32 + (σ3 − σ1 )2 l32 l12 . (2.20)

As can be seen from (2.20), τn2 is the positive value and turning to zero at the
principal sites. Indeed, if the normal ν coincides with one of the principal axes, then
one of the guide cosines equals to a unit, and the other two equal to zero, and then
τn2 = 0.
Now let’s look at octahedral sites. Since they are equally inclined to the principal
axes, all directional cosines for them are: li 2 = 1/3. We substitute these values l i 2 =
1/3 into formulas (2.19), (2.20) and obtain the values of normal and tangent stresses
on octahedral sites:
2.7 Maximal Tangent Stresses 23

σoct = (σ11 + σ22 + σ33 )/3 = (σ1 + σ2 + σ3 )/3,


τoct = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 /3 (2.21)

The formula for τoct through arbitrary coordinate components of stresses on the
basis of (2.21) is as follows:
/
τoct = (σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2 + 6(σ12
2
+ σ23
2
+ σ31
2
)/3.

Octahedral stresses are invariant values, which are expressed through stress
tensor invariants:
/
σoct = J1 /3, τoct = 2J12 − 2J2 /3.

These stresses were introduced by Nadai. They not only play an important role in
the calculation of equivalent stresses for a complex stress state, but also are used to
formulate plasticity criteria.
We will determine the sites on which maximal stress occurs. The position of
these sites can be determined by examining the extremum of the expression (2.20).
Because:

(σ1 − σ3 ) = (σ1 − σ2 ) + (σ2 − σ3 ),

and the square of the number is more than the sum of squares of numbers of its
components, so:

(σ1 − σ3 )2 ≥ (σ1 − σ2 )2 + (σ2 − σ3 )2

Therefore, τn reaches the maximum on those sites on which


/ the third of the terms
in formula (2.20) is maximal. Therefore, with l12 = l32 = 1 2, l22 = 0 we get that τn
= τmax (index 2 shows the axis which the site is parallel):
/
τmax = τ2 = (σ1 − σ3 ) 2 (2.22)

Thus, the site with maximal tangent stress is parallel to the principal axis 2 and is
equally inclined to the principal sites on which the maximal (σ1 ) and minimal (σ3 )
principal stresses act, (Fig. 2.9).
If you consider two more sites that are parallel to the first and third principal axes,
respectively, and to the other two are equally inclined, then we get two more extreme
values of tangent stresses:
/ /
τ1 = (σ2 − σ3 ) 2, τ3 = (σ1 − σ2 ) 2
24 2 Basic Concepts of Stress–Strain State Theory

Fig. 2.9 Site with maximal


tangent stresses

The values of τ1 , τ2 , τ3 are called the principal tangent stresses. For them, the
ratio is correct: τ1 + τ2 + τ3 = 0.
It should be noted that the surfaces on which the principal tangent stresses are
may not be mutually orthogonal. They form the sides of a regular dodecahedron. Its
sides are not free from normal stresses.
On sites where the principal tangent stresses act, normal stresses are:

σ2 + σ3 σ3 + σ1 σ1 + σ2
σn1 = , σn2 = , σn3 =
2 2 2

2.8 Stress Deviator and Spherical Part of Stress Tensor

A stress tensor exists at each point in the solid. Therefore, there is a field of stress
tensors in the body. The two properties of tensors that we will need in the future
should be noted:
• The sum of two tensors is a tensor, the components of which are the sum of the
corresponding components of the term tensors;
• The product of the tensor per scalar λ is a tensor whose components are λ times
larger than the corresponding components of the multiplied tensor.

Let’s consider the stress state, in which there are only three identical principal
stresses σ, equal to the average stress at a given point of the body, on three mutually
perpendicular sites:

σ1 = σ2 = σ3 = σ = σii /3 ≡ (σ11 + σ22 + σ33 )/3.

This stress state is described using the tensor:


⎛ ⎞
σ00
Tσ = ⎝ 0 σ 0 ⎠ (2.23)
00σ
2.8 Stress Deviator and Spherical Part of Stress Tensor 25

The tensor (2.23) is called a spherical stress tensor (the spherical part of a stress
tensor).
Subtract the spherical tensor from the stress tensor. As a result, we get a new
tensor, called a stress deviator.
⎛ ⎞
( ) σ11 − σ σ12 σ13
Dσ ≡ si j ≡ ⎝ σ21 σ22 − σ σ23 ⎠ (2.24)
σ31 σ32 σ33 − σ

Thus, the stress tensor at each point can be represented as the sum of the spherical
stress tensor (σ) and the stress deviator (sij ). For their components, such ratio is
performed:

σi j = si j + σδi j (i, j = 1, 2, 3), (2.25)

where δij is the Kronecker delta (2.4).


The decomposition of the stress tensor into the spherical and deviated parts is of
great importance when studying the behavior of elastic and plastic bodies under loads.
The spherical part extracts a uniform tension or compression from the stress state, in
which only the volume of this body element changes without changing its shape. The
stress deviator characterizes the shear state in which the shape of the element changes
without changing its volume. Therefore, the stress deviator indicates the deviation
of the stress state from the all-around tension (compression) or the deviation of the
deformed body shape from the original.
Some examples of the simplest stress states.
1. Uniform all-round tension (compression) of the body.
In this case, the stress deviator is zero and the spherical tensor is different from
zero. Vectors σ→n and n→ are collinear. The principal normal stresses equal to each
other.
2. Tension (compression) in one direction only.
In this case, the stress tensor is uniaxial, i.e., for example, σx x /= 0, and the
remaining components are all zero. If σx x > 0, then the stress tensor is the simple
tension tensor. If σx x < 0, then the stress tensor is the simple compression tensor.
If n→ is collinear with the x axis, then σ→n is also collinear with the same axis. If the
vector n→ is normal to the axis x, then σ→n = 0. The other two principal normal stresses
are identical to zero.
3. Plane stress state.
Let σzx = σzy = σzz = 0. If the vector n→ is collinear with the axis z, then σ→n = 0.
One of the principal stresses is zero, and the stress matrix has the form:
26 2 Basic Concepts of Stress–Strain State Theory
⎛ ⎞
σx x σx y 0
Ts = ⎝ σ yx σ yy 0 ⎠
0 0 0

Let’s build invariants for the spherical tensor and stress deviator.
The first invariant of the spherical tensor is the same as the first invariant of the
stress tensor:

J1s = 3σ ≡ σ11 + σ22 + σ33

The first invariant of the stress deviator is zero, since:

J1d = sii ≡ (σ11 − σ) + (σ22 − σ) + (σ33 − σ) = σ11 + σ22 + σ33 − 3σ = 0

We use the expression for the second invariant of the stress tensor (2.16), substi-
tuting instead of the stresses σii of the difference σii –σ, to determine the second
invariant of the stress deviator. After simple transformations, we get:
[
J2d = −si j si j /2 = − (σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2
( 2 )] [ ]
+6 σ12 + σ23 2
+ σ31
2
/6 = − (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 /6

The third invariant of the stress deviator has the form:

J3d = σi j σ jk σki /3

In the theory of plasticity, the concept of stress tensor intensity, which was
introduced by Ilyushin, is widely used.
Stress tensor intensity is formally defined through the second invariant of the
stress deviator:
( / )
σu2 = −3J2d = 3 2 si j si j

It follows that:
√ /
2
σu = (σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2 + 6(σ12
2
+ σ23
2
+ σ31
2
)
2
√ /
2
= (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 (2.26)
2
The stress tensor intensity is an invariant value, since it is expressed through the
second invariant of the stress deviator. It is exactly the same as the octahedral tangent
stress (2.21). The numerical coefficient in formula (2.26) is chosen so that in the case
of simple tension or compression (σ11 = σ1 , all other components equal to zero), the
condition σu = |σ1 | is fulfilled.
2.8 Stress Deviator and Spherical Part of Stress Tensor 27

Positive value:
√ /
√ 3
s= si j si j = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 , (2.27)
3
is called a stress deviator modulus.
The guide tensor of the stress deviator is called a tensor, each component is divided
into a modulus of the deviator (2.27):
⎛ ⎞
(σ11 − σ )/s σ12 /s σ13 /s

Dσ ≡ (s̄ i j ) = = ⎝ σ21 /s (σ22 − σ )/s σ23 /s ⎠ (2.28)
s
σ31 /s σ32 /s (σ33 − σ )/s

Since the first invariant of the deviator is zero, then: s ii = s 11 + s 22 + s 33 = 0. At


the same time: s i j s i j = 1. Therefore, the guide tensor is completely characterized
by four numbers, since its six components are already connected by the two given
relationships. It should be noted that the principal axes of the guide tensor coincide
with the principal axes of the stress tensor and the stress deviator tensor.
Using the definition of the guide tensor, the components of the stress tensor can
be represented in the following form:

σi j = σδi j + s s i j

Two modulus σ and s determine the scalar properties of the material. Its vector
properties are defined by a guide tensor ( s i j ).
Loading at a given point in the body is called simple, if all components of the
stress tensor change it in proportion to one common parameter λ, that is σi j = λσi0j .
In that case, the components of the guide tensor remained unchanged. In other cases,
loading is called complex.
A few words about an additional physical parameter—the Nadai-Lode param-
eter μσ , characterizing the stress state in three-dimensional deformable solids.
As it is known, the three-dimensional stress state can be reduced to several types
of the generalized stress state. This can be done using the Nadai-Lode parameter,
which characterizes the volumetric stress state in deformable bodies.
The Nadai-Lode parameter μσ is defined by this expression:

2σ2 − σ1 − σ3
μσ = , −1 ≤ μσ ≤ 1
σ1 − σ3

The Nadai-Lode parameter varies from “–1” to “1” and characterizes the type of
volumetric stress state:
• If μσ ∈ [−1; −0.5[—this corresponds to a generalized tension state;
• If μσ ∈ [−0.5; +0.5]—this corresponds to a generalized shearing state;
• If μσ ∈ ]0.5; +1]—this corresponds to a generalized compression state.
28 2 Basic Concepts of Stress–Strain State Theory

Using the parameter μσ , the stress tensor (2.17) can be decomposed into several
components. For example, the stress tensor (2.17) can be written as a sum of the
three simple stress states:
⎡ ⎤ ⎡ ⎤
( ) 100 10 0
σ1 + σ3 σ1 − σ3 ⎣ σ1 − σ3
TH = + μσ 0 1 0⎦ + (1 − μσ )⎣ 0 0 0 ⎦
2 2 2
001 0 0 −1
⎡ ⎤
00 0
σ1 − σ3 ⎣
+ 2μσ 0 0 0 ⎦,
2
0 0 −1

where the first tensor characterizes hydrostatic compression, the second is the
pure shear, and the third tensor is the uniaxial compression at μσ > 0 or uniaxial
tension at μσ < 0.

2.9 Strains and Displacements

Under the influence of external forces, all solids change their shapes (deformed).
Solid points change the position in space continuously. Solid does not contain breaks
and voids after deformation (that is, no cracks occur).
First, we define what is meant by the terms, “deformation” and “strain”.
Deformation indicates a change in solid geometry. This is a general concept for
various types of deformations: stretching, compression, twisting, bending, crushing,
etc.
Strain means a deformation measure, or the degree of deformation.
Consider, for example, a material sample with an initial length of L0. Stretch the
sample, that is, increase its length by a value of ΔL to the length L 1 L 0 + ΔL. Strain
is called dimensionless quantity ε:

ΔL L1 − L0
ε= =
L0 L0

It should be noted that strain is not additive. That is, if we add up the strains, then
the error accumulates.
For example, the first experimenter
/ stretched a sample with a length of 10 cm
by 1 cm. Strain equals to ε1 = 1 10. The sample length is now 11 cm. Another
experimenter does not know about/ the sample history and stretch it by another 1 cm.
The stain in this case is ε2 = 1 11. Then the total strain is:
/ /
ε = 1 10 + 1 11
2.9 Strains and Displacements 29

Fig. 2.10 To define a


displacement vector

However,/this is less than the strain obtained by this sample from its original
length ε = 2 10.
Now introduce the concept of “displacement” for a deformed solid.
A vector u→, having a start at point A of an undeformed body and an end at the
corresponding point A’ of the deformed body is called the full displacement vector
of the point, (Fig. 2.10). Its projections on the coordinate axis:

u→ = (u 1 , u 2 , u 3 ), u i ≡ u i (x), x ≡ (x1 , x2 , x3 )

The displacements thus defined for most deformed solids are small, compared to
the geometric sizes of the body.
It should be noted that, as a rule, we consider kinematically unchanged systems
in solid mechanics. That is, the movement of a body in space is not considered as a
rigid whole.
Based on the general statements of the theory of tensor field, the increment of the
displacement vector d u→ can be represented as:
[ ←
]/
d u→ = (E · d r→) + rot→
u×d r 2, (2.29)

where E is the symmetric strain tensor, the components of which are associated
with the components of the displacement vector u→ according to known formulas (see
below).
The second term in (2.29) describes the part of the point A’ displacement relative
to A that occurs when the small element containing point A rotates as a rigid body
around point A.
Thus, the “pure deformation” (strain) is described using the first term in the
expression (2.29), i.e., the symmetric tensor E.
The method of strain determining is that we calculate changes in the lengths of
linear elements, as well as changes in the angles between two linear elements by the
point displacements.
We get the formulas linking the point A displacement and the deformation in its
vicinity.
30 2 Basic Concepts of Stress–Strain State Theory

Fig. 2.11 To define Solid


strains

Let’s first consider a plane problem, that is, accept that:

u 3 (x) ≡ 0, u 1 ≡ u 1 (x1 , x2 ), u 2 ≡ u 2 (x1 , x2 )

After the deformation, segment AB with projections dx 1 , dx 2 occupies position


A’B’, (Fig. 2.11). The full displacement of point A along the axis x 1 is u1 . Point B
will move along with the point A by u1 plus additional displacement du1 due to the
deformation of segment AB along the axis x 1 .
Since the displacement increment is small and du1 = du1 (x 1 , x 2 ), so with accuracy
to terms of the higher order of small:

∂u 1 ∂u 1 ∂u 1
du 1 = dx1 + dx2 ≡ dx j ≡ u 1, j dx j ( j = 1, 2),
∂ x1 ∂ x2 ∂x j

where the first term corresponds to the component dx 1 extension along the axis
x 1 . The second term (since the deformations are small) describes the displacement
due to the rotation of the segment AB relative to the axis x 2 by an angle α1 :

∂u 1
α1 ≈ tgα1 = ≡ u 1 ,2
∂ x2

We consider the point B displacement along the axis x 2 , by analogy. As a result,


we get:

∂u 2 ∂u 2 ∂u 2
du 2 = dx1 + dx2 ≡ u 2, j dx j , α2 ≈ tgα2 = ≡ u 2,1 ( j = 1, 2).
∂ x1 ∂ x2 ∂ x1

Values describing linear extensions along axes are called linear strains and are
denoted by:

∂u 1 ∂u 2
ε11 = ≡ u 1,1 , ε22 = ≡ u 2 ,2
∂ x1 ∂ x2
2.9 Strains and Displacements 31

To describe shear strains, a value equal to half of the total change of right angle
among coordinate axes is used:
/ /
ε12 = (α1 + α2 ) 2 = (u 1 ,2 +u 2 ,1 ) 2

If the deformation of the neighborhood of point A is spatial (ui (x)/= 0, i = 1, 2,


3), then the increment of displacements will be:

du i = u i , j dx j (i, j = 1, 2, 3)

and such strains are added:


( ) ( )
ε33 = u 3,3 , ε32 = u 2,β + u 3,2 /2, ε31 = u 3,1 + u 1,3 /2

As a result, it is obtained that in the general case the small strains (u i2 , j << u i , j )
and displacements are connected by the Cauchy ratios:
( )
u i, j + u j,i
εi j = (i, j = 1, 2, 3). (2.30)
2
Remark 1. In the case of finite displacements (geometrically nonlinear theory of elas-
ticity), u i2 , j and u i , j are comparable. Then components of strain tensor are connected
with displacements by the following ratios:

( )
εi j = u i, j + u j,i + u k,i u k, j /2 (i, j, k = 1, 2, 3).

Remark 2. In the bending membranes and thin plate problems with transverse
loading, only deflections may be finite. Then only u 23 , j and u 3 , j are comparable.
The relations between strains and displacements take the form:

( )
εi j = u i , j + u j , i + u 3,i u 3 , j /2 (i, j = 1, 2, 3)

Strain components (2.30) form a tensor of the second rank E, which, like the stress
tensor, is symmetrical (εi j = ε ji ):
⎡ ⎤
εx x εx y εx z
E = ⎣ ε yx ε yy ε yz ⎦
εzx εzy εzz

This tensor fully describes the deformation of the vicinity of the considering point.
32 2 Basic Concepts of Stress–Strain State Theory

2.10 Principal Axes and Principal Values of Strain Tensor

Just as this was done when considering the stress tensor, it is possible to study in
which directions there are only linear strains and no shear strains. It is easy to show
that these directions also correspond to extreme values of strains. These extreme
values are called principal strains and they are denoted as: ε1 , ε2 , ε3 .
We can use operations to determine the principal directions and principal strains
when determining the principal directions and the principal values for the stress
tensor (see Sect. 2.6).
Thus, for strain tensor components, the following equation is true:
( )
εi j − εδi j l j = 0 (i, j = 1, 2, 3)

The solvability conditions of these equations are the characteristic equation for
the principal values of the strain tensor:

ε3 − ε2 I1 + εI2 − I3 = 0

The coefficients in this equation are three invariants, independent of the orientation
of the coordinate system:

I1 = εii ≡ ε11 + ε22 + ε33 = ε1 + ε2 + ε3 ,


I2 = (εii ε j j − εi j εi j )/2 = ε11 ε22 + ε22 ε33
+ ε33 ε11 − ε212 − ε223 − ε231 = ε1 ε2 + ε2 ε3 + ε3 ε1 ,
| |
| |
|| || | ε11 ε12 ε13 |
I3 = ||εi j || ≡ || ε21 ε22 ε23 || = ε1 ε2 ε3 (2.31)
|ε ε ε |
31 32 33

The strain tensor matrix in the principal axes is a diagonal matrix:


⎡ ⎤
ε1 0 0
E = ⎣ 0 ε2 0 ⎦
0 0 ε3

The first invariant I 1 has a simple geometric meaning that it is a relative volume
change:

ΔV − ΔV0
θ = u i ,i = lim = di vu
ΔV0 →0 ΔV0

The direction of the principal axes of the strain tensor is obtained from ratios
which are similar to (2.11), (2.12).
2.11 Strain Deviator and Spherical Part of Strain Tensor 33

It should be noted that within the framework of the theory of elasticity, the principal
axes of stress and strain tensors for isotropic solids coincide.

2.11 Strain Deviator and Spherical Part of Strain Tensor

Strain tensor (εij ) can be represented as the sum of two tensors:

εi j =εi j +εδi j (i, j = 1, 2, 3) (2.32)

/
where ε = θ 3 ≡ (ε11 + ε22 + ε33 )/3 is the spherical part of strain tensor; δij is
Cronk Characters (2.4); εij is the components of strain tensor deviator:
⎛ ⎞
( ) ε11 − ε ε12 ε13
Dε ≡ εij ≡ ⎝ ε21 ε22 − ε ε23 ⎠ (2.33)
ε31 ε32 ε33 − ε

The strain spherical tensor (ε) describes the volume deformation θ at the body
point. Its first invariant coincides with the first invariant of the strain tensor (2.31):

I1s = 3ε ≡ ε11 + ε22 + ε33

Strain deviator characterizes deformation of the shape change. Its first invariant
is zero, since:

I1d =εii ≡ (ε11 − ε) + (ε22 − ε) + (ε33 − ε) = ε11 + ε22 + ε33 − 3ε = 0

The second invariant of the strain deviator has the form (by analogy with the stress
deviator):

I2d = − εi j εi j /2
[ ]
= − (ε11 − ε22 )2 + (ε22 − ε33 )2 + (ε33 − ε11 )2 + 6(ε12
2
+ ε23
2
+ ε31
2
) /6
[ ]
= − (ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2 /6

The third invariant of the deviator has the following form:

I3d =εi j εjk εki /3

The second invariant of the strain deviator I 2d has a fundamental role in the theory
of plasticity.
The strain deviator modulus is a positive value:
34 2 Basic Concepts of Stress–Strain State Theory
√ /
√ 3
ε = εij εij = (ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2
3
The intensity of the strain tensor according to Ilyushin is determined through the
expressions for I 2d :

4 2
ε2u = − I2d = εi j εi j
3 3
or
√ /
2 ( )
εu = (ε11 − ε22 )2 + (ε22 − ε33 )2 + (ε33 − ε11 )2 + 6 ε212 + ε223 + ε231
√3 /
2
= (ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2 (2.34)
3
The geometric interpretation of the strain intensity is as follows: εu with accuracy
to the numerical multiplier coincides with the octahedral shift. Octahedral shift is the
relative shift of the site, which equals to the inclination to the principal strain axes.
In the case of simple tension or compression ε11 = ε1 , ε22 = ε33 = −νε1 ,
ε12 = ε23 = ε31 = 0 (ν is the Poisson’s ratio). Then from (2.34), we obtain:

2
εu = (1 + ν)|ε1 |
3
For incompressible material (ν = 0,5) εu = |ε1 |. Strain intensity according to (2.34)
is invariant.
A strain guide tensor is a deviator whose components equal to:

εi j =εi j /ε

It should be noted that: εii = 0 εi j εi j = 1/2 . Therefore, the strain guide tensor is
fully characterized by specifying four numbers, since the six components are already
connected by the two ratios. By using it, the components of the strain tensor can be
represented in the following form:

εi j = εεij + εδij

Two modulus ε and ε define the scalar properties and the guide tensor ( εi j ) define
the properties of material vector.
Deformation process is called simple if the values εi j = constant during
deformation. Otherwise, the deformation process is complex.
Strain rate tensor has the form:
2.12 Strain Compatibility Equations 35
⎡ ⎤
ε̇x x ε̇x y ε̇x z
Ė = ⎣ ε̇ yx ε̇ yy ε̇ yz ⎦
ε̇zx ε̇zy ε̇zz

2.12 Strain Compatibility Equations

The six independent components of the strain tensor are easily determined by the
known three differentiable components of the displacement field u i (x) using Cauchy
formula (2.30). The inverse operation is difficult, since not always six continuous
components εi j (x) correspond to any continuous displacement field. If a field exists,
then the strains are called compatible, otherwise—incompatible.
Compatible strains are related by relations, which are called strain compatibility
equations. They can be obtained in various ways.
On the one hand,(they can be)considered as integrable conditions of six differential
equations εi j (x) = u i, j + u j,i /2 relative to three unknown functions u i (x).
On the other hand, they express the physical fact that the whole body before and
after deformation is continuous and the particles of material are connected to each
other. These ratios were first obtained by Saint-Venan in 1860.
There is the procedure for obtaining the first group of compatibility equations,
which establishes links among strains acting on the same plane.
Write the Cauchy ratios in the components:

ε11 = u 1,1 , ε22 = u 2,2 , ε33 = u 3,3


( ) ( ) ( ) (2.35)
ε12 = u 1,2 + u 2,1 /2, ε23 = u 2,3 + u 3,2 /2, ε31 = u 3,1 + u 1,3 /2

We compose the expression, using (2.35) and the permutation property of the
differential order:
( )
ε11,22 + ε22,11 = u 1,122 + u 2,211 = u 1,2 + u 2,1 ,12 = 2ε12,12

We use the property that: 2ε12 = u 1,2 + u 2,1 .


By analogy, we get two other similar equations in other planes. As a result:

εαα,ββ + εββ,αα − 2εαβ,αβ = 0 (α, β = 1, 2, 3; α /= β) (2.36)

Recall that, as before, there is no summation for repeating Greek indices.


Let’s build a second group of three compatibility equations, which establishes
relationships among strains acting on different planes. To do this, we compose an
expression:

( ) ( )
ε12,3 − ε23,1 + ε31,2 ,1 = u 1,23 + u 2,13 − u 2,31 − u 3,21 + u 3,12 + u 1,32 ,1 /2 = u 1,231 = ε11,23
36 2 Basic Concepts of Stress–Strain State Theory
( )
From which: ε12,3 − ε23,1 + ε31,2 ,1 − ε11,23 = 0.
Similarly, two more similar equations can be obtained by using a circular index
permutation. As a result:
( )
εαα,γ − εβγ ,α + εγ α,β ,α − εαα,βγ = 0 (α, β, γ = 1, 2, 3; α /= β /= γ ) (2.37)

Thus, in order to find the corresponding displacement field on six continuous


components of the strain tensor, it is necessary to fulfill six differential equations of
compatibility (2.36), (2.37) in partial derivatives with respect to six components of
the strain tensor εij .
Remark.
In the case of a single-connected body, these equations are necessary and sufficient,
but they are only necessary for a multi-connected body.
Geometric meaning of compatibility relations. Imagine that the body before
deformation was divided into many material particles in the form of a rectangular
parallelepiped. Let us assume that each particle was subjected to an arbitrary strain
εij , after which the material particles took the form of coal parallelepipeds, which
may no longer constitute a solid. To prevent this, the strain components must satisfy
the Eqs. (2.36), (2.37).
It should be pointed out that when solving problems of the theory of elasticity in
displacements, the compatibility equations are satisfied automatically. When solving
the problem in stresses, the compatibility equations are among the basic equations
that must be satisfied.

Control Questions

1. What are the number of independent components of the strain tensor and stress
tensor?
2. What are the principal stress tensor sites?
3. At the principal sites of the stress tensor, the stresses are zeros:
1. normal;
2. tangent;
3. both.
4. In the main axes of the stress tensor, the stresses are extreme:
1. normal;
2. tangent;
3. octahedral.
5. Octahedral sites are equally inclined:
1. to coordinate axes;
2. to principal axes;
3. to the first and third principal axes.
2.12 Strain Compatibility Equations 37

6. Stress tensor components on inclined sites:


1. σνi = σi j l j
2. σi = σi j e j

1, i = j
3. δi j =
0, i /= j
7. Normal stress at octahedral sites:
/
1. σoct = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 3;
/ /
2. σoct = (σ11 + σ22 + σ33 ) 3 = (σ1 + σ2 + σ3 ) 3;
3. σoct = σi j e j .
8. Tangent stresses on octahedral sites:
/
1. τoct = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 3;
/
2. τoct = (σ1 − σ3 ) 2;
3. τ2oct = |σν |2 − σn2 .
9. Maximal tangent stresses:
/
1. τmax = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 3;
/
2. τmax = (σ1 − σ3 ) 2;

1, i = j
3. τmax =
0, i /= j
10. Tangent stresses are maximum at sites:
1. principal;
2. octahedral;
3. equally inclined to the 1st and 3rd principal axes.
11. Equilibrium equations are right:
1. inside the body;
2. on body surface;
3. everywhere.
12. Equilibrium equations:
1. S 3 − S 2 J1 + S J2 − J3 = 0;
2. σi j, j + ρFi = 0;
3. σi = σi j e j .
13. Boundary conditions are specified:
1. inside the body;
2. on body surface;
3. everywhere.
14. Boundary conditions in stresses:
38 2 Basic Concepts of Stress–Strain State Theory

1. σi j l j = Rνi ;
2. σνi = σi j l j ;
3. σ−ν = −σν .
15. What are absolute deformations and relative strains? Mathematical expressions
for absolute deformations and relative strains?
16. The invariants of Stress/strain tensor are retained during replacement:
1. coordinate systems;
2. loadings;
3. boundary conditions.
17. First, second, third invariants of stress tensor:
1. J = σ1 σ2 + σ2 σ3 + σ3 σ1 ;
2. J = σ1 σ2 σ3 ;
3. J = σ1 + σ2 + σ3 .
18. First, second, third invariants of strain tensor:
1. I = ε1 ε2 + ε2 ε3 + ε3 ε1 ;
2. I = ε1 ε2 ε3 ;
3. I = ε1 + ε2 + ε3 .
19. Linear strains in principal axes:
1. equal to zero;
2. are extreme;
3. coincide with angular.
20. Principal axes of stress and elastic strain tensors:
1. coincide;
2. mutually perpendicular;
3. make an angle 45° .
21. Stress intensity:
1. si j = σi j − σδi j ; /
2. σ = (σ11 + σ22 + σ33 ) 3;

3. σu = 2
2
(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 .
22. Deviatoric Stress tensor:
/
1. σ = (σ11 + σ22 + σ33 ) 3;
2. si j = σ√i j − σδi j ;
3. σu = 22 (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 .
23. Spherical part of stress tensor:
1. si j = σi j − σδi j ; /
2. σ = (σ11 + σ22 + σ33 ) 3;
2.12 Strain Compatibility Equations 39

3. σu = 2
2
(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 .
24. Spherical part of stress tensor characterizes:
1. uniform comprehensive tension (compression);
2. shift;
3. bend.
25. Deviatoric stress tensor characterizes:
1. uniform comprehensive tension (compression);
2. shift;
3. bend.
26. Deviatoric strain tensor:
1. I = ε1 ε2 + ε2 ε3 + ε3 ε1 ;
2. εi j = εi j − εδi j ; /
3. ε = (ε11 + ε22 + ε33 ) 3.
27. Spherical part of strain tensor:
1. I = ε1 ε2 + ε2 ε3 + ε3 ε1 ;
2. εi j = εi j − εδi j ; /
3. ε = (ε11 + ε22 + ε33 ) 3.
28. The intensity of strain tensor:
1. I = ε1 ε2 + ε2 ε3 + ε3 ε1 ;
2. εi j = ε√i j − εδi j ;
3. εu = 32 (ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2 .
29. In stresses/strains, the first invariant is zero:
1. deviator’s;
2. tensor’s;
3. spherical part of tensor.
30. Cauchy ratios for small strains:
/
1. εi j = (u i , j + u j ,i ) 2; /
2. γi j = (u i , j + u j ,i + u k ,i u k , j ) 2;

3. εu = 3
2
(ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2 .
31. Cauchy ratios bind:
1. stresses and strains;
2. strains and displacements;
3. principal axes of tensors.
32. Strain Compatibility Equations:
/
1. γi j = (u i , j + u j ,i + u k ,i u k , j ) 2;
40 2 Basic Concepts of Stress–Strain State Theory
( )
2. εαα,ββ + εββ,αα − 2εαβ,αβ = 0, εαα,γ − εβγ,α + εγα,β ,α − εαα,βγ = 0;

3. εu = 3
2
(ε1 − ε2 )2 + (ε2 − ε3 )2 + (ε3 − ε1 )2 .

Control Exercises

1. At a point P, the stress tensor relative to axes P x1 x2 x3 has components σi j . On


the area element d S (1) having the unit normal n̂1 , the stress vector is t(n̂1 ) , and
on area element d S (2) with normal n̂2 the stress vector is t(n̂2 ) .
Show that the component of t(n̂1 ) in the direction of n̂2 equals to the component of
t(n̂2 ) in the direction of n̂1 .
2. Verify the result established in Exercise 1 for the area elements having normal:

1( ) 1( )
n̂1 = 2ê1 + 3ê2 + 6ê3 and n̂2 = 3ê1 − 6ê2 + 2ê3
7 7
The stress matrix at P is given with respect to axes P x1 x2 x3 by:
⎡ ⎤
[ ] 35 0 21
σi j = ⎣ 0 49 0 ⎦
21 0 14

3. The stress tensor at P relative to axes P x1 x2 x3 has components in MPa given by


the matrix representation:

⎡ ⎤
[ ] σ11 2 1
σi j = ⎣ 2 0 2 ⎦,
1 20

where σ11 is unspecified.


Determine a direction n̂ at P for which the plane perpendicular to n̂ will be
stress-free, that is, for which t(n̂) = 0 on that plane.
What is the required
( value of σ11
) for this condition?
Answer: n̂ = 13 2ê1 − ê2 − 2ê3 , σ11 = 2 MPa.
4. The stress tensor has components at point P in ksi, as specified by the matrix:

⎡ ⎤
[ ] −9 3 −6
σi j = ⎣ 3 6 9 ⎦
−6 9 −6
2.12 Strain Compatibility Equations 41

Determine:
a) The stress vector on the plane at P whose normal vector is.

1( )
n̂ = ê1 + 4ê2 + 8ê3 ;
9

b) The magnitude of this stress vector;


c) The component of the stress vector in the direction of the normal;
d) The angle in degrees between the stress vector and the normal.
( ) √
Answer: a) t(n̂) = −5ê1 + 11ê2 − 2ê3 ; b) t(n̂) = 150; c) 23/9; d) 77.96°.
5. Let the stress tensor components at a point P be given by σi j = ±σ0 n i n j , where
σ0 is a positive constant.
Show that this represents a uniaxial state of stress having a magnitude ±σ0 and
acting in the direction of n i .
6. Show that the sum of squares of the magnitudes of the stress vectors on the
coordinate planes is independent of the orientation of the coordinate axes, that
is, show that the sum

(ê1 ) (ê1 ) (ê2 ) (ê2 ) (ê3 ) (ê3 )


ti ti + ti ti + ti ti

is an invariant.
7. With respect to axes Ox 1 x 2 x 3 the stress state is given in terms of the coordinates
by the matrix:

⎡ ⎤
[ ] x1 x2 x22 0
σi j = ⎣ x22 x2 x3 x32 ⎦
0 x33 x3 x1

Determine:
a) The body force components are functions of the coordinates if the equilibrium
equations are satisfied everywhere;
b) The stress vector at point P (1,2,3) on the plane whose outward unit normal
makes equal angles with the positive coordinate axes.
Answer:
( )
a) b1 = − 3xρ2 , b2 = − 3xρ3 , b3 = − xρ1 b) t(n̂) = √13 6e1 + 19e2 + 12e3 .
Λ Λ Λ

8. Relative to the Cartesian axes Ox 1 x 2 x 3 a stress field is given by the matrix:


42 2 Basic Concepts of Stress–Strain State Theory

Fig. E.8 Cylinder of radius


r and length L

⎡( ) ( ) ⎤
1 − x12 x2 + 23 x23 − 4 − x22 x1 0
[ ] ⎢ ( ) ⎥
− ( 2 3 2) (
x 3 −12x
σi j = ⎣ − 4 − x22 x1 0 ) ⎦
0 0 3 − x12 x2

a) Verify that in the absence of body forces the equilibrium equations are satisfied.
a) Show that the stress vector vanishes at all points on the curved surface of the
cylinder (Fig. E.8).

9. Rotated axes P x1' P x2' P x3' are obtained from axes P x1 x2 x3 by a righthanded
rotation about the line PQ that makes equal angles with respect to the P x1 x2 x3
axes (Fig. E.9).

Determine the primed stress components for the stress tensor in MPa.
⎡ ⎤
[ ] 30 6
σi j = ⎣ 0 0 0 ⎦,
6 0 −3

if the angle of rotation is:

a) 120°, or b) 60°.

Answer:
⎡ ⎤
[ ] 0 0 0
a) σi'j = ⎣ 0 −3 6 ⎦MPa,
0 6 3
2.12 Strain Compatibility Equations 43

Fig. E.9 Axis Q making


equal angles with x1 , x2 , x3

⎡ ⎤
[ ] −5 10 10
b) σi'j = 13 ⎣ 10 −11 −2 ⎦MPa,
10 −2 16

10. At the point P, rotated axes P x1' P x2' P x3' are related to the axes P x1 x2 x3 by the
transformation matrix:

⎡ √ √ ⎤
[ ] a√ 1 − 3 1 + √3
ai j = ⎣ 1 + 3 b√ 1 − 3 ⎦,

1− 3 1+ 3 c

where a, b and
[ c]are to be determined.
Determine σi'j if the stress matrix relative to axes P x1 x2 x3 is given in MPa by:

⎡ ⎤
[ ] 101
σi j = ⎣ 0 1 0 ⎦
101

Answer:
⎡ √ √ ⎤
[ ' ] 1 11 + 2√ 3 5 + 3 −1√
σi j = ⎣ 5 + 3 5√ 5 − √ 3 ⎦MPa
9
−1 5 − 3 11 − 2 3

11. At point P, the stress matrix is given in MPa with respect to axes P x1 x2 x3 by:
44 2 Basic Concepts of Stress–Strain State Theory
⎡ ⎤ ⎡ ⎤
[ ] 64 0 [ ] 211
case 1: σi j = ⎣ 4 6 0 ⎦, case 2: σi j = ⎣ 1 2 1 ⎦.
0 0 −2 112
Determine for each case:
a) The principal stress values;
b) The directions of principal stress directions.
Answer:
a) case 1: σ1 = 10 MPa, σ2 = 2 MPa, σ3 = −2 MPa;
case 2: σ1 = 4 MPa, σ2 = σ3 = 1 MPa.
b) case 1:
( )/ √ ( )/ √
n̂(1) = ± ê1 + ê2 2, n̂(2) = −ê1 + ê2 2, n̂(3) =
( )/ √
−ê1 − ê2 + 2ê3 6;
case 2: ( )/ √ ( )/ √
n̂(1) = ê1 + ê2 + ê3 3, n̂(2) = −ê1 + ê2 2, n̂(3) =
( /
) √
−ê1 − ê2 + 2ê3 6.

12. When referred to principal axes at P, the stress matrix in ksi is:

⎡ ⎤
[ ∗] 20 0
σi j = ⎣ 0 7 0 ⎦
0 0 12

If the transformation matrix between the principal axes and axes P x1 x2 x3 is:
⎡ 3 ⎤
[ ] − 5 1 − 45
1
ai j = √ ⎣ a21 a22 a23 ⎦,
2 − 3 −1 − 4
5 5
[ ]
where a21 , a22 , a23 are to be determined, calculate σi j .
Answer:
⎡ ⎤
[ ] 730
σi j = ⎣ 3 7 4 ⎦
047

13. The stress matrix in MPa when referred to axes P x1 x2 x3 is:


2.12 Strain Compatibility Equations 45
⎡ ⎤
[ ] 3 −10 0
σi j = ⎣ −10 0 30 ⎦
0 30 −27

Determine:
a) The principal stresses σ1 , σ2 , σ3 ;
b) The principal stress directions.
Answer:
a) σ1 = 23MPa, σ2 = 0MPa, σ3 = −47MPa
b) n̂(1) = −0.394ê1 + 0.788ê2 + 0.473ê3 ,
n̂(2) = 0.913ê1 + 0.274ê2 + 0.304ê3 ,
n̂(3) = 0.110ê1 + 0.551ê2 − 0.827ê3

14. At point P, the stress matrix relative to axes P x1 x2 x3 is given in MPa by:

⎡ ⎤
[ ] 5 a −a
σi j = ⎣ a 0 b ⎦,
−a b 0

where a and b are unspecified.


At the same point relative to axes P x1' x2' x3' the matrix is:
⎡ ⎤
[ ∗] σ1 0 0
σi j = ⎣ 0 2 0 ⎦
0 0 σ3

If the magnitude of the maximal shear stress at P is 5.5 MPa, determine σ1 , σ3 .


Answer: σ1 = 7 MPa, σ3 = −4 MPa.
15. The state of stress at point P is given in ksi with respect to axes P x1 x2 x3 by the
matrix:

⎡ ⎤
[ ] 10 2
σi j = ⎣ 0 1 0 ⎦
2 0 −2

Determine:
a) The principal stress values and principal stress directions at P;
b) The maximal shear stress value at P;
c) The normal n̂ = n i êi of the plane at P on which the maximal shear stress acts.
46 2 Basic Concepts of Stress–Strain State Theory

Answer:
a) σ1 = 2ksi, σ2 = 1ksi, σ3 = −3ksi;.
( )/ √ ( )/ √
n̂(1) = 2ê1 + ê3 5, n̂(2) = ê2 , n̂(3) = −ê1 + 2ê3 5;

b) τmax = ±2.5ksi;
( )/ √
c) n̂ = ê1 + 3ê3 10.
16. The stress tensor at P is given with respect to O x1 x2 x3 in matrix form with units
of MPa by:

⎡ ⎤
[ ] 4bb
σi j = ⎣ b 7 2 ⎦,
b24

where b is unspecified.
If σ3 = 3 MPa and σ1 = 2σ2 , determine:
a) The principal stress values;
b) The value of b;
c) The principal stress direction of σ2 .
Answer: a) σ1 = 8 MPa, σ2 = 4 MPa, σ3 = 3 MPa; b) b = 0, c) n̂(2) = ê1 .
17. The state of stress at P, when referred to axes P x1 x2 x3 is given in ksi by the
matrix:

⎡ ⎤
[ ] 93 0
σi j = ⎣3 9 0 ⎦
0 0 18

Determine:
a) The principal stress values at P;
b) The unit normal n̂∗ = n i êi∗ of the plane on which σn = 12 ksi and στ = 3 ksi.
Answer:
( √ )/ √
a) σ1 = 18 ksi, σ2 = 12 ksi, σ3 = 6 ksi; b) n̂∗ = ê1∗ + 6ê2∗ + ê3∗ 2 2.

18. Sketch the Mohr’s circles for the simple states of stress given by:
⎡ ⎤ ⎡ ⎤
[ ] σ0 0 σ0 [ ] σ0 0 σ0
a) σi j = ⎣ 0 σ0 0 ⎦, b) σi j = ⎣ 0 2σ0 0 ⎦.
σ0 0 σ0 σ0 0 −σ0
2.12 Strain Compatibility Equations 47

Determine the maximal shear stress in each case.


Answer: a) τmax = σ0 , b) τmax = 1.5σ0 .
19. Relative to axes O x1 x2 x3 , the state of stress at O is represented by the matrix:

⎡ ⎤
[ ] 6 −3 0
σi j = ⎣ −3 6 0 ⎦
0 0 0
⎡ ⎤
[ ] 300
Show that, relative to principal axes σi j = ⎣ 0 9 0 ⎦, the stress matrix is:
000
⎡ ⎤
[ ] 300
σi j = ⎣ 0 9 0 ⎦
000

Show that, these axes result from a rotation of 45°about the x3 axis.
Verify these results by Equations:

σ11 + σ22 σ11 − σ22


σ' 11 = + cos 2θ + σ12 sin 2θ,
2 2
σ11 + σ22 σ11 − σ22
σ' 22 = − cos 2θ − σ12 sin 2θ,
2 2
σ11 − σ22
σ' 12 = sin 2θ + σ12 cos 2θ
2

20. The representation of stress matrix at P is given by:

⎡ ⎤
[ ] 29 0 0
σi j = ⎣ 0 −26 6 ⎦
0 6 9

Decompose this matrix into its spherical and deviator parts, and determine the
principal deviator stress values.
Answer: s1 = 25 MPa, s2 = 6 MPa, s3 = −31 MPa.
21. At point P in a continuous body, the stress tensor components are given in MPa
with respect to axes P x1 x2 x3 by the matrix
48 2 Basic Concepts of Stress–Strain State Theory
⎡ √ ⎤
[ ] 1 −3 √2
σi j = ⎣ −3 1 − 2 ⎦
√ √
2− 2 4

Determine:
a) The principal stress values σ1 , σ2 , σ3 together with the corresponding principal
stress directions;
b) The stress invariants Ji , i = 1, 3;
c) The maximal shear stress value and the normal to the plane on which it acts;
d) The principal deviator stress values;
Answer:
a) σ1 = 6MPa, σ2 = 2MPa, σ3 = −2MPa;
( √ )/ ( √ )/
n̂(1) = ê1 − ê2 + 2ê3 2, n̂(2) = ê1 − ê2 − 2ê3 2, n̂(3) =
( )/ √
ê1 + ê2 2;

b) J1 = 6, J2 = −4, (J3 = −24;


√ √
Λ √ ) √ Λ Λ

c) τmax = 4MPa, n̂max = (1 + 2)e1 − (1 − 2)e2 + 2e3 /2 2;


d) s1 = 4MPa, s2 = 0(MPa)2 , s3 = −4(MPa)3 .
22. In a solid body, the stress field relative to axes O x1 x2 x3 is given by:

⎡ ( ) ⎤
(x1 x2 2 ) ( x31 1 − x)/
2 2
[ ] 2 0
σi j = ⎣ x1 1 − x2 x2 − 3x2 3 0 ⎦

2 0 2x32

Determine:
a) The body force distribution if the equilibrium equations are satisfied throughout
the field; ( √ )
b) The principal stresses at P a, 0, 2 a ;
c) The maximal shear stress at P;
d) The principal deviatoric stresses at P.
Answer:
a) b1 = b2 = 0, b3 = −4x3 /ρ;
b) σ1 = 8a, σ2 = a, σ3 = −a;
c) τmax = ±4.5a;
d) s1 = 16a/3, s2 = −5a/3, s3 = −11a/3.
2.12 Strain Compatibility Equations 49

23. Let the stress tensor components σi j be derivable from the symmetric tensor
field ϕi j by the equation: σi j = εiqk ε j pm ϕkm,q p .
Show that, in the absence of body forces, the equilibrium equations are satisfied.
Chapter 3
Material and Solid Mechanical
Characteristics (Properties)

3.1 Main Mechanical Characteristics of Materials


and Solids

Mechanical properties of materials characterize the material behavior in various


mechanical force fields. Mechanical properties are important in the study of defor-
mation processes, since they predefine the characteristics of the stress-strain state
(SSS) of body under certain boundary conditions.
The general mathematical model for describing solid mechanical behavior
contains physical equations between the stress-strain state components. These equa-
tions include parameters or characteristics, and the functions of material mechanical
properties in some cases.
For nonhomogeneous materials, as a rule, using standard test methods, the
material mechanical properties, the sample mechanical properties, and the solid
mechanical properties are different and significant in some cases.
For materials with relative structural homogeneity and small sizes of struc-
tural elements (for example, metals), using standard test methods, the mechan-
ical properties of the samples practically correspond to the material mechanical
properties.
The mechanical properties of the materials obtained during the sample tests are
relative, since the values of the same properties may differ under other test conditions.
It must be clearly understood that physical and mechanical values characterize the
properties of a particular sample of material. It can be shown that the material samples
used to determine its mechanical properties must not be smaller than the size of the
elementary volume.
The material mechanical properties can be divided into several groups:
a) Strength properties—characterize the ultimate material resistance to various
types of loads;
b) Elastic properties—characterize material elastic deformability under loads;
c) Dynamic properties—characterize the conditions of elastic vibrations, the
transmission and propagation of strain waves and stress waves in materials;
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 51
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_3
52 3 Material and Solid Mechanical Characteristics (Properties)

d) Rheological properties—characterize the material deformation in time at given


loading conditions.
Let’s look at the selected groups of material mechanical properties.
Strength properties
Most often, properties such as compressive and tensile strengths, bonding, coefficient
(angle) of internal friction are used to characterize the material strength properties.
/
The ultimate strength σ∗ is the stress when the sample breaks: σ ∗ = P F,
where P is the crushing load, F is the applied load area.
The ultimate strength at uniaxial compression of material samples, or compres-
sive strength σ pr ess , is most often the determined and used characterization of the
strength of the material.
Typically, the higher the compressive strength of the materials, the higher their
density.
The value of the material compression strength determined on the standard sample
for many materials is most often taken as the main basic characteristic with which
other mechanical characteristics are connected.
Tensile strength σ p is an important indicator of the mechanical properties of
materials.
Remark. Carrying out experiments to determine the value σ p by direct stretching
for a large class of materials is very difficult for many reasons. Therefore, tensile
strength is often recommended to be calculated.
Importantly, the tensile strength σ p for a large range of materials is significantly
lower than their compressive strength. For example, it is one of the characteristic
features of rocks, concrete, and other materials. At the same time, for most metals,
the compressive and tensile strengths are practically same.
Shear strength can be characterized by two functionally related parameters:
bonding and internal friction angle (coefficient) of the material. This functional
relationship is expressed most often by the Coulomb1 ) equation:

τn = σn tgϕ + τ0 , (3.1)

where τn is the shear stress (tangent stress); σn is the normal shear stress; ϕ is the
internal friction angle (respectively tgϕ is the internal friction coefficient); τ0 is the
cohesion.
The cohesion τ0 (C) represents the limit of shear resistance on a site where
there is no normal pressure, i.e., no resistance to shear effort due to internal friction,
(Fig. 3.1).
The internal friction angle ϕ or internal friction coefficient tgϕ characterizes
the growth of shear stress intensity with an increase of normal stresses, i.e., it is a

1 Charles-Augustin de Coulomb (1736–1806), French military engineer and physicist. He is best


known as the eponymous discoverer of what is now called Coulomb’s law, the description of the
electrostatic force of attraction and repulsion, though he also did important work on friction.
3.1 Main Mechanical Characteristics of Materials and Solids 53

Fig. 3.1 Illustrative diagram for explaining the physical meaning of the “cohesion” characteristic

Fig. 3.2 Illustrative diagram for explaining the physical meaning of the “internal friction angle
(coefficient)” characteristic

proportional coefficient between increments of tangent dτn and normal dσn stresses
at shearing, (Fig. 3.2):

τn 2 − τn 1 dτn
tgϕ = = .
σn 2 − σn 1 dσn

The greater the actual normal forces, the greater the tangent forces to move the
body, (Fig. 3.2).
Importantly, although the Coulomb Eq. (3.1) is used for the irreversible strain
zones, it can also be used for elastic zones if the contact displacements in them are
large enough.
Equation (3.1) is often written in principal stresses:

σ1 − (2λ + 1)σ3 = σ pr ess , (3.2)

where σ1 / and σ3 are respectively maximal and minimal principal normal stresses;
λ = sin ϕ (1 − sin ϕ); σ pr ess is the ultimate strength at uniaxial compression.
A feature of the strength criterion (3.1) or (3.2) is that this criterion takes into
account the destruction both as a result of shear and breakaway. Shear fracture occurs
when the tangent stress acting on the shear plane and depending on the normal stress
of this plane reaches a value determining the strength of the material. Breakaway
fracture occurs when the minimal normal stress reaches the material strength at
uniaxial compression.
Elastic properties of materials
Elastic properties are characterized by modulus of elasticity E at uniaxial stress state
(modulus of longitudinal elasticity or, otherwise, Young2 ) modulus); shear modulus

2Thomas Young (1773–1829), a British polymath who made notable contributions to the fields of
vision, light, solid mechanics, energy, physiology, language, musical harmony, and Egyptology.
54 3 Material and Solid Mechanical Characteristics (Properties)

μ (G); k is volume expansion coefficient; transverse strain coefficient (Poisson3 )


coefficient)).
Modulus of elasticity E is the ratio of normal stress σn to relative linear elastic
deformation of the sample εl = Δl/l in the direction of the applied loading action
(i.e., the ratio of stress to strain under uniaxial compression or tension):
/
E = σn εl .

The shear modulus μ (G) is the ratio of the tangent stress to the relative shear
(i.e., the ratio of the tangent stress to the shear strain, for example, when there is a
torsional sample):
/
μ = τ θ.

Relative shear θ is called angular strain. It characterizes the shape change of


deformable body and is expressed by the following formula:
/
π 2−α
θ= / ,
π 2

where α is inclined angle of each rectangular body element after deformation.


The volume expansion coefficient (or volumetric modulus of elasticity or all-
round compression modulus) equals to the ratio of uniform all-round stress to the
sample volume relative elastic change:
σv
k= ,
ΔV /V
/
where ΔV V is the relative volume change.
The transverse strain coefficient ν, or Poisson’s ratio, is a measure of the propor-
tionality between relative deformations in the direction perpendicular to the applied
load vector and parallel to the one (ratio of transverse strains to longitudinal strains
with sample under simple loading):

Δd/d εd
ν= = .
Δl/l εl

Elastic properties are functionally connected. The relationship between the basic
physical constants for an ideal elastic body is shown in Table 3.1.
Remark. Lamé constants are often called constants λ and μ.

3Baron Siméon Denis Poisson (1781–1840), a French mathematician, engineer, and physicist who
made many scientific advances.
3.1 Main Mechanical Characteristics of Materials and Solids 55

Table 3.1 Relationship between basic physical constants for an ideal elastic body
Basic constants Constants pairs
λ, μ μ, ν E,ν E,μ k, μ
2μν νE μ(E−2μ)
λ λ 1−2ν (1+ν)(1−2ν) 3μ−E k − 23 μ
μ μ μ E
2(1+ν) μ μ
k λ + 23 μ – E
3(1−2ν) – k
(3λ+2μ)μ 9kμ
E λ+μ 2(1 + ν)μ E E 3k+μ
λ (E−2μ) 3k−2μ
ν 2(λ+μ) ν ν (2μ) 6k+2μ

Thus, if we know any two of these characteristics, then we can calculate the values
of the others. Typically, the properties of E and ν are experimentally determined using
experimental samples.
As the density of materials increases, their modulus of elasticity tends to increase.
Poisson’s ratio (transverse deformation coefficient) ν of materials theoretically varies
within such limits −1 ≤ ν ≤ 0.5.
The Poisson’s ratio for most “normal” materials has positive value in the range
(0.2–0.5). However, the theoretical values of ν for an isotropic material are in the
range of −1 ≤ ν ≤ 0.5. The coefficient ν can’t exceed 0.5, as such materials would
increase their density during extension and would become unstable.
Materials with negative Poisson’s ratio are called auxetic. These mate-
rials are capable to extend/compress in the direction perpendicular to uniaxial
extending/compression, respectively.
The possibility of the existence of such materials is confirmed by the known
relation of the isotropic elasticity theory (see table 3.1):

3k − 2μ
ν= . (3.3)
6k + 2μ

It can be obtained from (3.3) that negative values ν are possible when performing
such condition:
( / )
μ > 3 2 k,

i.e., when the shear modulus is more than 50% of the volume strain modulus.
Remark. All known theoretical and experimental approaches to create composites
with ν < 0 can be divided into two categories: either all components of the composite
have ν > 0 or some of them are auxetic.
Elastic properties together with material density ρ define dynamic properties of
material—the ability of distribution of elastic strain waves or elastic vibrations in
materials.
Dynamic properties are characterized by the propagation velocity C, acoustic
resistance Q and absorption coefficient α of elastic waves.
56 3 Material and Solid Mechanical Characteristics (Properties)

Among the various types of elastic oscillation in solids, the most interesting are
longitudinal waves (waves of compaction-rarefaction, P-waves), transverse waves
(shear strain propagation waves, S-waves) and surface waves (Rayleigh waves).
In the longitudinal waves, the direction of particle oscillation coincides with
the direction of wave propagation. In transverse waves, the direction of particle
oscillation is perpendicular to the direction of wave propagation. Surface waves are
oscillations of the solids’ surface.
Longitudinal wave propagation velocity:
/ /
λ + 2μ E(1 − ν)
CP = = .
ρ ρ(1 + ν)(1 − 2ν)

Transverse wave propagation velocity:

/ /
μ E
CS = = .
ρ 2ρ(1 + ν)

Surface wave propagation velocity:

C R = Kν CS ,

where K ν is the dimensionless coefficient depending on Poisson’s ratio (for example,


at ν = 0.25, K ν = 0.9194; at ν = 0.5, K ν = 0.9553).
The propagation velocities of these waves are characterized by the following
inequality:

C P > CS > C R .

The product of the material density by the velocity of the corresponding wave is
called the acoustic resistance or acoustic rigidity.

Q = ρ · C.

The acoustic resistance characterizes the effect of the medium properties on the
intensity I (frequency) of oscillations in this medium.
The velocity of longitudinal elastic waves is the most common characteristic. The
velocity of elastic waves in materials increases with the increase of compressive loads.
For materials having a small elastic limit, elastic waves are acoustic oscillations. Two
types of waves are possible in a limitless medium: longitudinal and transverse waves.
The absorption of mechanical energy is impossible in an ideal elastic body, only the
dissipation of mechanical energy. The mechanical characteristics of most materials
depend on the speed at which the load acts.
The strength and yield limits of most materials are several times higher than that
in static tests with a small number of dynamic loading cycles. At the same time,
3.2 Classification of Materials by the Nature of Deformation 57

with multiple loading with a load varying in value and in sign, destruction can occur
at stresses much lower than static strength. With a high-speed dynamic loading (for
example, blow-up), the failure mechanism has a different character from considering
these processes from the point of view of “classical representations”. In this case,
there are complex wave processes: first, there is the material hardening and then
(with repeated periodic loads) there is the material fatigue destruction.
It should be noted that the dynamic modulus of elasticity is less dependent on the
experimental conditions than the static modulus of elasticity. Therefore, the dynamic
modulus of elasticity more reflects the properties of the samples as a material,
wherein the dynamic modulus of elasticity increases as the samples increase and
approaches the calculated effective modulus.

3.2 Classification of Materials by the Nature


of Deformation

Materials are divided into several basic types depending on their response to external
influences.
First of all, a large group is composed of materials that behave as elastic media
under the loading action. A perfectly elastic material is characterized by direct
proportionality between the applied loads and the deformations.
A large number of materials exhibit rheological properties that characterize the
growth of deformations (absolute and relative deformations) in time under constant
external loads (creep) or a gradual stress decrease under constant deformations
supported by constant loads (stress relaxation). Stress relaxation is a reverse creep
process. During relaxation, elastic strains in the material gradually change over time
into plastic ones, but the total strains do not change with time. The stresses decrease
in this case.
The strength and elasticity of natural materials under the prolonged action of
large loads are reduced, asymptotically approaching some limit values—the limit of
long-term strength σ∞ and the long-term limit of modulus of elasticity of long-term
elasticity E ∞ . For most materials: σ∞ = (0.7 − 0.8)σ; E ∞ = (0.65–0.95)E.
When stresses are greater than the elastic limit, both elastic and plastic strains
appear. The plastic strains remain after the body is not loaded.
The plastic properties can be characterized by a plastic coefficient, which is
the ratio of the work W pl spent on breaking a given material volume to the work
Wel needed to break the same material volume with the same value of compressive
strength value under the assumption of the material ideal elasticity:
/
K pl = W pl Wel .
58 3 Material and Solid Mechanical Characteristics (Properties)

Such characteristic as brittleness is important for a large class of materials. Brit-


tleness is the material ability to destroy under the influence of applied loads without
significant residual (plastic) strains.
Brittleness can be characterized by a brittleness coefficient, which is the ratio of
the work Wel taken to deform a material sample to the elastic limit to the total work
W f r to destroy a sample:
/
K f r = Wel W f r .

A characteristic property of most natural materials is their high degree of hetero-


geneity and anisotropy. A material is called homogeneous material if its physical
properties are the same at all the points. A material is called isotropic material with
respect to a property, if this property at a point is the same in all directions. If
the properties of a material at a point depending on the direction, then it is called
anisotropic material.
Real materials can be considered quasi-isotropic in many applications.

3.3 Complete Diagrams of Deformation and Fracture

It seems important, based on the nature of the process under consideration, to establish
the fact of which position is preferable to study the process: from the position of
deformations or stress states.
The difference between the specified loading and specified deformation modes
can be illustrated in the following diagrams, (Fig. 3.4).
If the sample is loaded with a load of weight P, then such a mode can be called a
specified loading mode (Fig. 3.3a). In this case, the displacements of the sample do
not affect the acting load, but on the contrary, the displacements are a direct result
of the acting load.
The most characteristic example of this loading mode is the effect of external
loads on the object, for example, a beam under the load.
If the sample resistance R with external influence on it practically does not affect
the displacements Δl1 on the sample surface, then this interaction is called specified
deformation mode (Fig. 3.3b). This deformation mode is typical when the load on
the sample is large stress arising from external influence.
A typical example of this type of effect on an object is external pressure through a
screw press. Press movement Δl1 determines the degree of influence on the sample.
In the mutual deformation mode (Fig. 3.3c), the contact surface displacement
of the sample with the external load depends on the “reactive resistance” of the
sample and on the external load, or the external load depends on the sample surface
displacement.
This loading mode is typical, for example, for a double-layer beam. The beam
is loaded by compressive loads on the right and left sides. In this case, additional
pressure is exerted on the lower layer as a consequence of the bending of the upper
3.3 Complete Diagrams of Deformation and Fracture 59

Fig. 3.3 Diagrams of specified loading and deformation modes. a—specified loading mode on
sample (R = P, Δl = f 1 (P), Δl does not affect P); b—mutual deformation mode on sample (R
almost does not affect Δl 1 ); c—interaction deformation mode (Δl = f 1 (Q, R)); d—combined
deformation mode (Δl = f 1 (P, Q, R))

Fig. 3.4 Scheme of the complete rock deformation diagram

layer. But at the same time, the underlying layer exerts opposition (pressure) on the
upper one.
In practice, cases of a combined mode of “operation” are very common, when
at the same time there are mutual deformation mode and specified loading mode
(Fig. 3.3d).
It should be noted that traditional tests of rock samples for the specified loading
mode often do not correspond to the real mechanical state of rocks in the massif,
where the given deformation mode mainly takes place. In traditional methods for
testing building materials and in real building structures, as a rule, the material
“works” in the given loading mode.
60 3 Material and Solid Mechanical Characteristics (Properties)

The methods of testing materials (especially natural materials), developed in


recent years, allow you to build complete deformation diagrams.
These diagrams characterize the mechanical state of the sample in all loading
areas: up to strength (pre-limit branch of the diagram); at the level of strength; beyond
the strength, including the section of residual strength (beyond the limit branch of
the diagram).
It should be noted that complete sample deformation diagrams reflect, as a
rule, “instant” mechanical states and do not characterize the material rheological
properties.
Studies of the material behavior and properties at the pre-limit stage of defor-
mation were carried out quite a lot. Similar studies for the “over-limit deformation
section” are single. However, the practical importance and relevance of these studies
are evident. For example, in an over-limit deformation branch, the solid takes on a
block structure.
Let’s consider the complete deformation diagram using the example of a rock
sample deformation. The deformation diagram under the uniaxial compression
becomes more informative when it is supplemented by the graphical dependence
of compressive stresses not only on longitudinal strains, but also on transverse
strains, (Fig. 3.4). We consider the case of uniaxial compression. We use the
following symbols: σ1 is compressive normal stress; e1 is longitudinal principal
linear compression strain; e3 is transverse principal linear strain, having negative
values.
With the increase of the external compressive load to the level σ1a , there are
processes of closing microcracks and pores, which is accompanied by the increase
of the effective modulus of elasticity E e f to the value of the modulus of the material
elasticity E, i.e., in this section, the dependence σ1 (e1 ) is non-linear. The relation-
ship between σ1 (e3 ) and the strain level e3a is also nonlinear, but | to /a lesser
| extent
than σ1 (e1 ). At this stage, the transverse strain coefficient β = |Δe3 Δe1 | (where
Δe
| 1 , Δe
/ 3 are| increments of corresponding strains) increases to a constant value
|Δe3a Δe1a |. The volume of the sample decreases as a result of its compression.
From the level of compressive stresses σ1a to the level σ1b , there is an elastic
compression of the sample with a constant modulus of elasticity E. Experimental
relationships between σ|1 (e1 )/and σ|1 (e3 ) are linear at this stage. The transverse defor-
mation coefficient β = |Δe3 Δe1 | < 0, 5 and remains constant up to the strain e1b
and e3b , that is, it represents the Poisson’s ratio at this stage. The volume of the
sample continues to decrease.
At a stress level greater than σ1b , the formation of microcracks begins. Thus,
for example, the first step of the sample breaking may begin to be realized. In this
case, dependence σ1 (e1 ) remains linear, and dependence σ1 (e3 ) (due to the formation
and opening of vertical microcracks) becomes non-linear. As compressive stress σ1
increases, the increase rate of transverse strain e3 is faster than that of longitudinal
strain e1 , i.e., the coefficient of transverse strain β increases, and the decrease in
sample volume slows down. If the external load does not increase in the future, then
the formed microcracks will stop their growth, and when the load is removed, they
will close.
3.3 Complete Diagrams of Deformation and Fracture 61

When the load is more than σ1c , then the process of unstable microcrack develop-
ment begins. First, it is the most dangerous, and then, as the external load increases,
and all others. During this period, the second fracture stage is realized. It should
be emphasized, that at stresses greater than σ1c , the crack development has an
unstable avalanche-like character. The development of the most dangerous microc-
rack weakens the loaded sections of sample and increases the acting stresses in the
vicinity of other cracks, causing their development. As a result, main macrocracks
are formed, leading to the sample destruction. The duration of the avalanche-like
cracking process depends on the level of active stresses in the range from σ1c to σ1d .
Stress level σ1c is insufficient to implement the fracture process, that is, the sample
destruction will not occur even after an infinitely long period of time. The stress level
σ1c can be interpreted as the ultimate long-term strength of materials for uniaxial
∞ ∞
compression σ pr ess . Since stresses are greater than σ pr ess , destruction will occur after
a certain period of time, the value of which decreases with increasing stresses, while
at stress σ1d , destruction occurs almost instantly. Stress level σ1d is interpreted as the
instant strength of materials for uniaxial compression σ pr ess .
In the stress interval from σ1c to σ1d , dependence σ1 (e1 ), as well as dependence
σ1 (e3 ), become non-linear, that is, the modulus of elasticity E becomes variable, and
E(e1 ) has the physical meaning of the deformation modulus. With stresses increasing
from σ1c to σ1d , the value E(e1 ) decreases from E to zero. In the interval cd, the
transverse strain |(e3 )| grows faster than the longitudinal one e1 , and the transverse
strain coefficient β(e1 ) increases. Starting with the stress level σ1c (corresponding

to the long-term strength σ pr ess ), for most natural materials there is a tendency to
increase their deformable volume (dilatancy phenomenon).
The maximum of sample loading capacity σ1d (equals to σ pr ess ), depends on
the experimental loading mode. Therefore, in the specified deformation mode, the
value σ1d is slightly larger than that in the specified loading mode, which is usually
realized on traditional experimental equipment. The specified deformation mode is
often referred to as a “rigid” loading mode, and the specified loading mode is called
“soft” loading mode.
At the maximum of loading capacity, in case of soft loading mode, main macro-
scopical cracks are formed in the sample and its destruction occurs (third stage of
destruction). In this case, the bearing capacity of the sample drops to zero without
increasing strains, i.e., the beyond-limit branch of the deformation diagram is practi-
cally vertical. At the same time, the destruction is often dynamic with the separation
of parts of the destroyed sample. Thus, with a soft loading mode, we only fix the
maximum of sample bearing capacity and not fix the material ultimate strength.
With a rigid loading mode, the maximum of sample bearing capacity σ1d corre-
sponds to the limit of material instantaneous strength for uniaxial compression σ pr ess .
Furthermore, in the section de (beyond the limit branch of diagram), the increase
of strain corresponds to a decrease in the bearing capacity of the sample σ1 to a
certain minimal value σ1e , called the material residual strength σost . Slope of diagram
beyond-limit branch σ1 (e1 ) is characterized by decline modulus, numerical value of
which is determined by inclination angle arctgM of beyond-limit branch σ1 (e1 ) to
negative direction of the axis e1 .
62 3 Material and Solid Mechanical Characteristics (Properties)

With a rigid loading mode, successive steps of the sample breaking can be possibly
observed at this stage, which is practically impossible to do with a soft loading mode
when the breaking is dynamic.
At the beyond-limit deformation section de, the bearing capacity of the sample is
accompanied by an increase in the sample transverse strain |e3 |, causing an increase
in the transverse strain coefficient β to values greater than one.

Control Questions

1. What is the difference between the Mechanical properties of samples and


mechanical properties of the material itself?
2. What are the main groups of mechanical properties of materials?
3. What are the basic strength properties of materials?
4. What are the basic elastic properties of materials?
5. What are auxetics (auxetic materials)?
6. What are the dynamic material properties?
7. What is the material classification according to the nature of the “response” to
external mechanical loading? What are the terms of “rheology” and “plasticity”?
8. What are isotropy and anisotropy with respect to some mechanical properties?
9. What are the various modes of mechanical loading of samples (specified loading
mode and specified deformation mode). Give some examples.
10. What is the complete diagrams of deformation and fracture at uniaxial
compression and tension?
11. What are the definitions of the modulus of elasticity and modulus of deforma-
tion?
Chapter 4
Construction of Mathematical Model
Problems

4.1 The System of Equations to Describe the Medium


Stress–Strain State

The mathematical condition of the physical (in-situ) existence of a solid is the


continuity equation:

∂ρ
+ div(ρ→
v ) = 0, (4.1)
∂t

where ρ is the medium density; v→ is the medium velocity.


Remark. div(ρ→ v ) = ∂∂x ρvx + ∂∂y ρv y + ∂∂z ρvz .
The differential equilibrium equations of a continuous medium in stresses
σi j , j +ρ Fi = 0 are performed at any continuous strain of all continuous media.
However, various real deformable bodies behave differently under the same
external loads. Thus, these equations with the addition of appropriate boundary
conditions are not sufficient to describe the specific solid motion.
Such system of equations is not complete. Therefore, to determine the stresses,
strains, and displacements of a solid under the external loads, equilibrium equations,
boundary conditions, strain compatibility equations and Cauchy equations are not
enough. To construct a closed and complete system of resolving equations, it is
still necessary to have physical relations between stresses and strains, taking
into account the behavior peculiarities of materials.
It should be noted that for a large number of tasks, it seems appropriate and
justified to use the hypothesis of small strains. This hypothesis allows you to limit
yourself to the geometrically linear setting of problems, that is, to exclude second
and higher degrees of strains from the equations, which greatly simplifies analytical
studies. However, the hypothesis of small strains does not contradict reality, given
that the displacements in most cases are negligible compared to the size of the bodies
themselves. Saint-Venan principle is important for a wide class of solid mechanics

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 63
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_4
64 4 Construction of Mathematical Model Problems

problems. This principle is that the forces applied to the small surface of the body
produced only local stresses and strains.
Thus, a complete system of equations constructed by a mechanical-mathematical
model describing the states of a solid mechanical object includes:
• Differential equilibrium equations of a continuous medium;
• Boundary conditions;
• Physical relations between stress and strain states of solid;
• Strain compatibility equations.

4.2 Physical Relationships Determining the Solid Behavior

Since a huge number of different materials are available, many physical laws
describing their behavior under various loading conditions are also possible. Unfortu-
nately, it was impossible to establish a universal physical law describing the behavior
of all materials. Therefore, we seek to establish relationships that can describe the
most important types and behaviors of materials under certain loadings and deforma-
tion conditions. As a result, a specific mathematical model is introduced. Model type
is significantly determined by experimentally established relations between stress
and strain states at points of deformable medium. The validity and correctness of the
model describing the real material state and behavior are tested experimentally.
Structural models are often used to describe the solid behavior and build math-
ematical models based on them. A large number of structural models are known.
Examples of the most famous structural models are given in Fig. 4.1. Among them,
for example: the elastic Hooke model (Fig. 4.1a), the viscous model (Fig. 4.1b),
the viscous-elastic relaxing model (Fig. 4.1c), the viscous-elastic Kelvin-Foigt
model (linear elastic-late) (Fig. 4.1d), the Hohenemser-Prager model (Fig. 4.1e),
Point-Thompson model (Fig. 4.1i), rigid-plastic model (Fig. 4.1g), elastic–plastic
model (Fig. 4.1j), viscoelastic model of Shvedova-Bingam (Fig. 4.1h), model with
the connection of serial elements (Fig. 4.1k), model with elastic delay and stress
relaxation (Fig. 4.1l) etc.
Thus, the process of material loading behavior can be mathematically described
using the abstract representation of deformable materials in the form of some struc-
tural models consisting of elementary structural units, each of these units is a certain
element (elastic, plastic or viscous).
Structural units with elastic properties are similar to springs with the Hooke
deformation law:

σ = E · ε,

where σ is stress; ε is strain; E is modulus of elasticity (proportionality coefficient).


Structural units with viscous properties correspond to the Newton’s deformation
law:
4.2 Physical Relationships Determining the Solid Behavior 65

а b

d
c

i
g

Fig. 4.1 Examples of some structural models


66 4 Construction of Mathematical Model Problems

Fig. 4.1 (continued)

σ = η · (dε/dt),

where dε/dt is the strain rate; η is the viscosity coefficient.


Remark. They are represented by a damper with holes in a cylinder containing
viscous liquid.
The plastic properties of structural units are modeled by dry friction. In this case,
deformation can occur only at stresses exceeding a certain value σm called the yield
strength. The rigid-plastic element can be characterized by cohesion (C) and internal
friction angle (coefficient) (tgϕ).
The deformation of materials in accordance with different laws can be reflected by
an appropriate combination of different elements in the structural model. Typically,
the views of structural model and parameters of model elements are set according to
special experiments, obtaining a family of curves corresponding to different levels
of active stresses.
There are some principal differences among elastic, plastic, and rheological
models of solid behavior.
• The state of elastic body after the load change is uniquely determined by the new
conditions and does not depend on the previous one.
• Plastic models “remember” the history of previous loads. But since the change in
states occurs instantly, in plastic models, this story is not connected with time in
any way.
• If the plasticity of materials characterizes their behavior at stresses exceeding the
limit of elasticity, creep is also come through at stresses less than the limit of
elasticity, but under the influence of loads for a long time.
Consider some of the defining points regarding the description of the behavior
of deformable solids within different approaches and models.
• Elastic Medium Model
4.2 Physical Relationships Determining the Solid Behavior 67

Based on the elastic model, it is assumed that some so-called “natural” states of
the medium can be distinguished, in which the stress tensor is identical to zero. For
all other states of elastic medium at any point and at any time, the stress tensor is a
one-to-one function of the strain tensor.
It should be noted that the stress state of the elastic medium at a given time only
depends on its deformed state, not on the way of transition from the natural state to
the current state and the speed of this way.
In the linear theory of elasticity, it is assumed that each point makes small displace-
ments relative to some “natural” states. If the medium is homogeneous and isotropic,
then there is a classical theory of elasticity which is based on the linear connection
between stresses and strains.
The relationship between stresses and strains is described using the Hooke’s law.
• Elastic–plastic Model
Common in applications is the theory of small elastoplastic deformations. This
theory belongs to the number of deformation theories of plasticity. Plastic deforma-
tion occurs when stresses reach the required limit value and is irreversible. In case of
elastic–plastic deformation, the strain increment is composed of two parts: elasticity
and plasticity: dei j = de(i j) + de[i j] . Square brackets in the indices denote the values
related to the plastic region, and parentheses refer to elastic.
It is assumed that the elastic and plastic properties of the material are manifested
independently. Plastic (residual) deformations occur when stresses reach a certain
limit value. For ideal plastic bodies, there are (some ) fixed relationships among stress
components called the plasticity condition: ϕ σi j = constant. Elastic–plastic prob-
lems usually include the cases where elastic and plastic strains occurring in the body
are of the same order.
The setting of elastic-plastic problems should take into account the ratios in the
elastic and plastic zones and at the border between them:
• Elastic region:
– Relationship between stresses and strains in the form of Hooke’s law;
– Boundary conditions formulated with respect to stresses or/and displacements;
– Compatibility conditions;
– Equilibrium equations.
• Plasticity region:
– Relationship between stresses and strain increments;
– Compatibility conditions of complete deformations;
– Plasticity condition;
– Boundary conditions at the boundary of the plastic region;
– Equilibrium equations in stresses.
• Boundary between elastic and plastic regions:
– Stresses and displacements change continuously, i.e., coupling conditions must
be carrying out at the boundary of elastic and plastic zones.
68 4 Construction of Mathematical Model Problems

• Viscous-Elastic Model

Viscoelastic media are solids whose deformation changes with time after applying
constant external loadings. If in such medium, the relations between stresses and
strains are described using linear differential operators (relative to time), then it is
called a linear viscous-elastic medium.
Rheological models are used to simulate viscoelastic processes. Rheological
models have the memory of previous medium states.
Rheological structural models can be represented by different connections of
elastic springs and viscous Newtonian bodies (a piston moving in a cylinder with
viscous liquid).

Control Questions

1. What is the definition of the Saint-Venan principle of loading static equivalence?


2. What is the complete system of equations for constructing a mechanical-
mathematical model that describes the stress–strain state of a solid?
3. What are the structural models for describing the solid behavior? Give some
examples.
4. What is the structural element of elastic medium? Explain the Hooke’s law.
5. What is the structural element with viscous properties? Explain the Newton’s
law.
6. What is the structural element with plastic properties? Explain the element “dry
friction”.
7. What is the principal difference among elastic, plastic, and rheological models
of solid behavior?
8. What are the defining positions regarding the description of the deformable
solid behavior within elasticity model?
9. What are the defining positions regarding the description of the deformable
solid behavior within elastic–plastic model?
10. What are the defining positions regarding the description of the deformable
solid behavior within viscous-elastic model?
Chapter 5
Mathematical Models of the Theory
of Elasticity

5.1 Basic Concepts and Definitions

The theory of elasticity is a part of the continuum mechanics and the solid
mechanics. The theory of elasticity deals with the determination of a stress–strain
state in elastic bodies at specified external loads.
The following basic hypotheses and assumptions regarding material properties,
loads and the nature of deformations are accepted in the theory of elasticity:
• Homogeneity and continuity hypothesis—the assumption makes it possible to
study the mechanical properties of bodies on samples with relatively small sizes
and allows you to use a differential calculus apparatus to study strain states;
• Assumption of small strains—at the points on the body, strains are taken so small
that they do not significantly affect the mutual location of loads applied to the body,
which allows us to consider the geometry of body unchanged when constructing
the equilibrium equations;
• The Euler and Cauchy stress principle—in each cross section, mentally drawn
inside the body, there is a force interaction according to the type of loads distributed
on the body surface, that is, using the method of sections, the effect of one part of
the body on another can be replaced by the surface forces acting on the section.
• The axiom of solidification (freezing)—at any fixed time t, the material body is
considered as an absolute solid, and the laws of theoretical mechanics, including
Newton’s laws, are valid for it, while the axiomatic concepts of force and mass
are used.
In establishing the constitutive relations of the theory of elasticity, in addition to
the accepted ones, the following hypotheses are introduced:
• The material ideal elasticity hypothesis—the ideal elasticity is the ability of
body to restore its original shape and size after “eliminating the causes” that
caused its deformation (removal of external loads, temperature, electromagnetic
and radiation fields, etc.);

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 69
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_5
70 5 Mathematical Models of the Theory of Elasticity

• Assumption of linear dependence between deformations and loads (Hooke’s


law)—it is assumed that for most materials, the displacements resulting from
the body strains are directly proportional to the stresses that caused them.
We assume that in its initial state, there is no stress in the body, the object has a
constant temperature and is in thermodynamic equilibrium. Therefore, in the initial
(“natural”) state of the elastic body, the stress tensor is identical to zero. For all other
states of elastic body at any point and at any time, the stress tensor is a one-to-one
function of the strain tensor.
It should be noted that the stress state of the elastic medium at a given time depends
only on its deformed state, not on the method of transition from the natural state to
the studied one and the speed of this transition.
In the linear theory of elasticity, it is assumed that each point makes small
displacements relative to some natural states. If the medium is also homogeneous
and isotropic, then we come to the classical theory of elasticity.
The classical theory of elasticity is based on a linear connection between stresses
and strains and is a linearized theory of the behavior of elastic media in the vicinity
of a natural state, when for any point of media, the displacements, and their gradients
relative to some natural states are small.
Thus, the stress at each point of the elastic body is a unique strain function:

σi j = ϕi j (εkm ). (5.1)

The loading path/deformation path, respectively, is the process of the stress/strain


tensor changing, depending on some monotonically increasing parameters (which
we call “time”).
We emphasize that, in fact, real time does not play a role in determining the model
of elastic body. Using this term, we talk only about the sequence of events, not about
the time of their action.
The stress tensor or strain tensor can be represented by vectors whose components
equals to the components of the corresponding tensors:

σ = σ(σ11 , σ22 , σ33 , σ12 , σ23 , σ31 ), ε = ε(ε11 , ε22 , ε33 , ε12 , ε23 , ε31 ).

Thus, the loading or deformation paths can be represented as curves described


using the ends of the vectors σ and ε in the respective spaces.
The law of elasticity [i.e., Eq. (5.1)] establishes, in particular, that the closed
deformation path corresponds to the closed loading path, and vice versa.
The specific potential energy of deformation is the value defined as the work
performed during the deformation of unit volume body:

∫ εi j
U= σi j dεi j .
εi j =0
5.1 Basic Concepts and Definitions 71

The internal energy change (due to the symmetry of the tensors εi j and σi j ) can
be written in an expanded form:

dU = σ11 dε11 + σ22 dε22 + σ33 dε33 +


+ 2(σ12 dε12 + σ23 dε23 + σ31 dε31 ). (5.2)

For a perfectly elastic body, the work done on an elementary volume in a closed
cycle by strain or stress is zero. The condition equals to zero work on an arbitrary
closed cycle can be represented as:
∫ ∫
dU = σi j dεi j = 0.

In order to perform the calculated ratio, it is necessary that the sub-integral


expression is a complete differential, i.e.,

∂U
σi j = . (5.3)
∂εi j
( )
The specific potential strain energy U εi j is a unique strain function, also called
the elastic potential energy. However, on the other hand, the expression:

dU ∗ = ε11 dσ11 + ε22 dσ22 + ε33 dσ33


+ 2(ε12 dσ12 + ε23 dσ23 + ε31 dσ31 )

is also a complete differential, since:

dU + dU ∗ = d(σi j εi j ).

For this reason:


∂U ∗
εi j = . (5.4)
∂σi j

The value U ∗ (σi j ) is called the additional specific strain energy and is an elastic
stress potential.
For the introduced energies U (εi j ) and U ∗ (σi j ), the general ratios are correct:

∂U ∂U ∗
dU (ε) = dεi j , dU ∗ (σ) = dσi j .
∂εi j ∂σi j

The meaning of the U and U * values for a nonlinear elastic material can be clearly
explained using the example of a uniaxial stress state, (Fig. 5.1).
The figure shows that the values:
72 5 Mathematical Models of the Theory of Elasticity

Fig. 5.1 Visual


representation for the
concepts of additional
specific strain energy and
elastic stress potential

∫ ε ∫ σ
U (ε) = σdε, U ∗ (σ) = εdσ
0 0

complement each other to the rectangle “σε” under the curve and above curve of the
“stress–strain” diagram.
It should be emphasized that the assumption of the existence of specific potential
strain energy does not necessarily mean a linear relationship between stresses and
strains. Linearity is introduced only by Hooke’s law.

5.2 Hooke’s Law

The relationship between stresses and deformations for an elastic body is described
using the Hooke’s law.1
The generalized Hooke’s law in the tensor representation is as follows:

1+ν νΣ 1
e= TH − I, (5.5)
E E
where e is the principal strain tensor, corresponding to principal stresses σi ; T H is
the stress tensor; Σ 1 is the first invariant of stress tensor T H ; I is the unit tensor.
Remark The relationship (5.5) equals to three vectors:

σi νΣ 1
ei = (1 + ν) − .
E E

1The same law was independently opened in 1680 by the French physicist E. Mariott (Mariotte E.,
1620–1684).
5.2 Hooke’s Law 73

The tensors in (5.5) can be written in any coordinate system. Constants E and
v fully characterize ideal elastic medium. However, you can also specify two other
constants. It is important that there is a relationship between each pair of constants
(see Chap. 3).

The generalized ratios of Hooke’s law for isotropic bodies are as follows:

1 1
ε11 = [σ11 − ν(σ22 + σ33 )], ε22 = [σ22 − ν(σ33 + σ11 )],
E E
1
ε33 = [σ33 − ν(σ11 + σ22 )],
E
1 1 1
ε12 = σ12 , ε23 = σ23 , ε31 = σ31 , (5.6)
2G 2G 2G
where E and G are, respectively, modulus of elasticity and shear modulus, v is
Poisson’s ratio. They are related by a known dependency: 2G = E/(1 + ν).
In solving the problems of the theory of elasticity, there is a need for inverse
relations when stresses are expressed through strains. Using the ratios (5.6), we get:

σi j = 2μεi j + λθδi j , (5.7)

where:
1 − 2ν
θ = ε11 + ε22 + ε33 = (σ11 + σ22 + σ33 ), (5.8)
E
θ is the volume strain; λ and μ are Lamé constants that are associated with the
dependencies of G, v and E (see Chap. 3):

νE E λ
λ= , μ=G= , ν= .
(1 + ν)(1 − 2ν) 2(1 + ν) 2(λ + μ)

Recall that for an ideal elastic body from five elastic constants λ, μ, K, ν, E, only
any two are independent. The dimensionalities of the values λ, μ, K, E correspond
to the stress or pressure dimensions. These constants are positive. Poisson’s ratio is
dimensionless. Its change interval is −1 ≤ ν ≤ 1/2. Value ν = 1/2 corresponds to
incompressible material. For all known isotropic materials, we have ν > 0.
Equation (5.7) with respect to εij is the inverse of Hooke’s law:
[ ]
1 3ν
εi j = σi j − σδ i j . (5.9)
2G 1+ν

With ν → 0.5 Lamé constant λ → ∞, which, according to the expression (5.8),


corresponds to incompressible material (θ = 0). In this case, it is difficult to use the
ratio (5.7). Therefore, it is recommended to record two separate ratios in which the
volume deformation would be explicitly written out. This is achieved, for example,
74 5 Mathematical Models of the Theory of Elasticity

by using the expression of the Hooke’s law through the spherical and deviatoric
components of stress and strain tensors:

si j = 2G C- i j , σ = K θ, (5.10)

2 E
K =λ+ μ= ,
3 3(1 − 2ν)

where K is the volume strain modulus.


For incompressible materials, instead of the second in Eq. (5.10), this condition
is used: θ = 0. The value θ is defined by (5.8).
It follows from Hooke’s law in the form (5.10), that the specific potential strain
energy can be divided into two independent parts:
( )
1 1 σ2 1 ε2
Ud = si j C- i j = σi j σi j − , Uv = σθ = K , (5.11)
2 4G 3 2 2

where U d is called the specific energy of the form change; U v is called the specific
energy of the volume change.
In the expanded form, formula (5.11) takes the following form:

1 − 2ν
Uν = (σ11 + σ22 + σ33 )2 ,
6E
1+ν( )
Ud = (σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2
6E
1 ( 2 )
+ σ + σ232
+ σ312
.
2G 12
Both values U d and U v play the important roles, for example, in formulating
strength criteria and flow laws during plastic deformation.
The relation (5.7) can be generalized to the case of an arbitrary anisotropic
material, assuming a linear relationship between the stress tensor and strain tensor
components in the form of:

σi j = E i jkl εkl . (5.12)

The relation (5.12) contains, in general, nine equations having nine terms each.
The components of E ijkl are coordinate independent values for homogeneous mate-
rial. Since stresses and strains depend on the orientation of coordinate system, elastic
constants E ijkl must also correspond to this dependence. They form 81 elastic constant
matrices and are transformed into components of a fourth-rank tensor, which is called
the elastic modulus tensor. Elastic constants do not depend on the orientation of
coordinate system only for isotropic materials.
In view of the symmetry of tensors σi j and εkl , the elastic modulus tensor is also
symmetrical with respect to the indices i and j, k and l: E ijkl = E jikl = E ijlk = E jilk .
5.2 Hooke’s Law 75

Thus, the independent components remain 36. The matrix of its components is as
follows:
⎡ ⎤
E 1111 E 1122 E 1133 E 1112 E 1113 E 1123
⎢E E 2223 ⎥
⎢ 2211 E 2222 E 2233 E 2212 E 2213 ⎥
⎢ ⎥
⎢ E 3311 E 3322 E 3333 E 3312 E 3313 E 3323 ⎥
⎢ ⎥.
⎢ E 1211 E 1222 E 1233 E 1212 E 1213 E 1223 ⎥
⎢ ⎥
⎣ E 1311 E 1322 E 1333 E 1312 E 1313 E 1323 ⎦
E 2311 E 2322 E 2333 E 2312 E 2313 E 2323

A further decrease in the number of independent components is obtained from


thermodynamic considerations, assuming the existence of a specific potential strain
energy. Substituting the ratios (5.12) and (5.3) into the expression for the specific
potential energy difference (5.2) in turn, we have:

∂U
dU = σi j dεi j = E i jkl εkl dεi j , dU = σi j dεi j = dεi j ,
∂εi j

where
∂U
= E i jkl εkl .
∂εi j

After re-differentiation and taking into account the possibility of changing the
order of differentiation, we obtain:
( ) ( )
∂ ∂U ∂ ∂U
= E i jkl , = E kli j .
∂εkl ∂εi j ∂εi j ∂εkl

Therefore,

∂ 2U
= E i jkl = E kli j .
∂εi j ∂εkl

Therefore, the number of independent constants for the anisotropic body becomes
21. If the elastic body has a plane of symmetrical elastic properties, then for it, the
number of independent constants is reduced to 13. For a body having three mutually
orthogonal symmetry planes (orthotropic body), the number of constants is reduced
to 9. The number of independent constants for an isotropic body, as previously shown,
is two.
The generalized Hooke’s law can be reversed by expressing strains through
stresses. Then:

εi j = Di jkl σkl . (5.13)


76 5 Mathematical Models of the Theory of Elasticity

The fourth-rank tensor Dijkl (elastic pliability tensor) has the same symmetric
properties as the elastic modulus tensor E ijkl .
It should be emphasized that the number of elastic constants in Hooke’s law
is reduced only when the symmetric planes are taken as coordinate. Other coordi-
nate systems would still contain 21 constants, expressed through nine independent
constants.

5.3 Clapeyron’s Formula and Clapeyron’s Theorem

Clapeyron’s formula
We use the expression (5.2) to calculate the specific potential strain energy. We
consider an arbitrary element of the volume of an elastic body, the stress–strain state
of which is given by the values σi j and εi j . We introduce a local system of principal
axes for stresses and strains. Let the principal stresses receive increments that are the
same as their final values (this is proportional loading). Then the work performed,
assigned to the unit of volume, is defined as:

W = (σ1 ε1 + σ2 ε2 + σ3 ε3 )/2.

It is independent for the loading path, and according to the law of conservation
of energy, it equals to the specific potential strain energy U. In arbitrary coordinate
axes, accordingly, we obtain:

1
U= σi j εi j . (5.14)
2
This expression is called the Clapeyron’s formula. We can write it in an expanded
form:
1
U= (σ11 ε11 + σ22 ε22 + σ33 ε33 ) + σ12 ε12 + σ23 ε23 + σ31 ε31 .
2
Taking into account the generalized Hooke’s law in form (5.12) or (5.13), the
Clapeyron’s formula (5.14) takes the form:

1 1
U= E i jkl εi j εkl or U = Di jkl σi j σkl .
2 2
Hence:
∂U 1 ∂ 1
= (Dmnkl σmn σkl ) = Dmnkl (δim δ jn σkl + δik δ jl σmn )
∂σi j 2 ∂σi j 2
1
= (Di jkl σkl + Dmni j σmn ) = Di jkl σkl ,
2
5.3 Clapeyron’s Formula and Clapeyron’s Theorem 77

or
∂U
= εi j . (5.15)
∂σi j

It should be noted that this ratio should not be lumped together with a very similar
ratio (5.4), which instead of U contains U * . At the same time, the ratio (5.15) is
assumed to satisfy the Hooke’s law and therefore the linear-elastic behavior of the
material. Meanwhile, the ratio (5.4) follows the general thermodynamic laws.
Remark Formulas (5.4) and (5.15) are partial formulas of Castigliano’s2 theorem
(see below).
Clapeyron’s theorem
Let the linearly elastic body under the influence of body forces F i and surface forces
Ri be in an equilibrium state. Let’s scalarly multiply the equilibrium equations by
the displacements ui and then integrate on the body volume V (with the surface S).
As a result, we obtain:
∫ ∫
u i σi j , j dV + u i ρFi dV = 0.
V V

We transform the sub-integral function in the first integral using the formula:

(u i σi j ), j = u i , j σi j + u i σi j , j ,

and, therefore:
∫ ∫ ∫
(u i σi j ), j dV + u i ρFi dV = u i , j σi j dV .
V V V

The first integral on the left side can be transformed into a surface integral using
the known Gauss3 —Ostrogradsky formula. In the right part, due to the symmetric
stress tensor, equality will be performed: u i , j σi j = εi j σi j . As a result, we get the
ratio:
∫ ∫ ∫
u i σi j l j dS + u i ρFi dV = εi j σi j dV .
S V V

The sub-integral expression in the right part according to the Clapeyron’s formula
(5.14) equals to twice the specific potential strain energy. Therefore:

2 Castigliano C. A. (1847–1884), Italian scientist-mechanics.


3 Gauss C. F. (1777–1855), German mathematician, physicist; works in the field of algebra, number
theory, differential geometry.
78 5 Mathematical Models of the Theory of Elasticity
∫ ∫ ∫
u i Ri dS + u i ρFi dV = 2 U dV . (5.16)
S V V

Thus, the work of elastic strain equals to half of the work performed by the external
surface and body forces on the displacements from the initial state of equilibrium to
the final state (Clapeyron’s theorem).

5.4 Thermoelasticity

We assume that the body is in a heterogeneous and non-stationary temperature field


T (x i , t). Initial temperature value is T 0 . Complete linear strains εii consists of strains
from the force load εii' and temperature expansion εii'' = αT (Neumann4 hypothesis):

εii = εii' + εii'' .

Temperature increments for shear strains are zeros. Therefore, the Hooke’s law,
taking into account the temperature, has the form:
[ ]
1 3ν
εi j = σi j − σδ i j + αT δi j . (5.17)
2G 1+ν

The value α is called the coefficient of linear thermal expansion of the material.
By expressing stresses through strains from (5.17), we obtain:
[ ]
νθ 1+ν
σi j = 2G εi j + δi j − αT δi j . (5.18)
1 − 2ν 1 − 2ν

The ratio between the deviators in thermoelasticity remains the same as in the
case of ideal elasticity (5.10). Only the spherical parts connected by stress and strain
tensors take another form:

si j = 2G C- i j , σ = K (θ − 3αT ).

Strains in formulas (5.17) and (5.18) are related to displacements by Cauchy ratios
(2.35).
For complex thermoelastic problems, in general, it is also previously necessary
to solve the heat distribution problem.
The temperature T (xi , t) at each point of the body must satisfy the thermal
conductivity equation

4 Franz Ernst Neumann (1798–1895), German physicist.


5.5 The Boundary Problems of Theory of Elasticity 79

∂T Q
= BΔ T + , (5.19)
∂t cγ

where B is the coefficient of heat transfer of materials; Δ is the Laplacian; Q(xi , t)


is the function showing the heat quantity produced by the thermal energy source in
unit volume and per unit time; c is the specific heat capacity; γ is the specific weight.
The initial condition of temperature distribution T (xi , 0) = f (xi ) should be added
to Eq. (5.19). In addition, the conditions of heat exchange with the environment (three
types of heat exchange conditions) shall be executed at the solid boundary.
The Q value in Eq. (5.19) for solids generally depends on the stress–strain state.
In this case, they talk about coupled thermoelastic problems.
For an elastic anisotropic body, Hooke’s law (5.12) takes the form:

σi j = E i jkl (εkl − αkl T ),

where αkl is a coefficient of the thermal expansion tensor of materials.


Constant E ijkl is determined under isothermal conditions at T = T 0 . If the temper-
ature increment is not small, then E ijkl and αkl should be considered as functions of
temperature.

5.5 The Boundary Problems of Theory of Elasticity

For the solution of the mathematical theory of direct problems in elasticity, that is
for definition of 15 unknown functions ui , σij , εij (i, j = 1, 2, 3), the following system
of the equations is available:
• Equilibrium equations

σi j , j +ρFi = 0, (5.20)

• Cauchy ratios

1
εi j = (u i , j +u j ,i ), (5.21)
2
• Hooke’s Law

σi j = 2μεi j + λθ δ i j . (5.22)

The system (5.20)–(5.22) includes 15 linear equations in partial derivatives.


When displacements are not explicitly unknown, Eq. (5.21) is replaced by the
conditions of strain compatibility (without summation by repeating indices):

εαα ,ββ +εββ ,αα −2εαβ ,αβ = 0,


80 5 Mathematical Models of the Theory of Elasticity

(εαβ ,γ −εβγ ,α +εγα ,β ),α −εαα ,βγ = 0. (5.23)

To close a resolving system of equations, you must add conditions on the boundary
(boundary conditions).
The most common situation is that the surface of the body S can be represented
as consisting of two parts: S = S σ + S u . It is assumed that the boundary conditions
at each point of surface are as follows:

u i = u i0 (x) on Su , Rνi = σi j l j on Sσ . (5.24)

This simplest case is not the only one and does not exhaust all possible options
for boundary conditions. It will be appreciated that other combinations of boundary
conditions are possible. Note that if the body contains an infinitely distant point, the
requirement of the regularity of the solution at infinity should to be added to the
boundary conditions, which, as a rule, is reduced to the constraint condition. There-
fore, in the general case, the theory of elastic problem is to solve Eqs. (5.20)–(5.23)
under boundary conditions (5.24).
It is also possible to set a theory of elastic inverse problem. In this case, you specify
stresses, strains, or displacements for all internal points of the body as coordinate
functions. Stress–strain state at the body’s boundary (boundary conditions) must be
defined.

5.6 The Theory of Elastic Boundary Problems


in Displacements

Depending on the setting of the boundary problem, the main defined functions are
either stresses or displacements.
If we substitute into equilibrium Eq. (5.20) the Hooke’s law (5.22) and exclude
strains using Cauchy relations (5.21), we get a system of three differential equations
with three unknown functions ui (Lamé equation):

(λ + μ)θ,i +μΔ u i + ρFi = 0, (5.25)

where θ = u1,1 + u2,2 + u3,3 = ui,i is the volume strain; Δ is the Laplacian.
The system of Eq. (5.25) is elliptical by Petrovsky5 in the area V body with all
values of the Poisson’s ratio, except for ν = 0.5 and ν = 1.
To construct a complete system of resolving equations, boundary conditions
should be attached to the three Eq. (5.25).
On one part S u of the boundary surface, displacements u0i can be set as known
coordinate functions:

5Petrovsky Ivan Georgievich (1901–1973), Russian mathematician, rector of Moscow State


University (1951), academician of the USSR Academy of Sciences.
5.6 The Theory of Elastic Boundary Problems in Displacements 81

u i = u i0 (x) on Su . (5.26)

External surface forces Ri = Ri (x i ) can be set on the part Sσ of the boundary


surface. Moreover, since the problem is solved in displacements, the values Ri should
be associated with the values ui using the Hooke’s law and Cauchy ratios. As a result,
the boundary conditions in displacements take the form on the part Sσ :

λθδi j l j + μ(u i , j +u j ,i )l j = Ri . (5.27)

In general, three types of boundary problems in displacements occur.


The first type of boundary problems consists in determining the displacements
and stresses within the elastic body if the displacements of the point are known on
the body surface (S = S u ).
The mathematical setting of the first boundary problem is as follows. Determine
the elastostatic state [u, σ] of the medium D corresponding to the mass force X and
satisfy the Lamé equation:

μΔ u + (λ + μ)grad divu + ρX = 0

by the boundary condition:


( )
∀y ∈ S : lim u ± (x) = f (y) x ∈ D± , (5.28)
x→y

where μ, λ are Lamé constants; f = ( f 1 , f 2 f 3 ) is the specified displacement vector


at the body boundary; S is the boundary of domain D.
If the domain contains an infinitely point, then the solution must satisfy the
conditions of regularity at infinity:
( )
lim |u(x)| = 0, lim ∂u j /∂ xi R = 0, (i, j = 1, ..., m), (m = 2, 3),
|x|→∞ R→∞

where R is the distance from current point x to the origin.


In the second boundary problem, the distribution of forces on the surface (S = S σ )
is known. The mathematical setting of the second boundary problem is as follows.
Determine the elastostatic state [u, σ] of the medium D corresponding to the mass
force X and satisfy the Lamé Eq. (5.25) by boundary condition:
( )
∀ y ∈ S lim (T(∂ x, n x )u(x))± = f (y) x ∈ D ± , (5.29)
x→y

where T(∂ x, n x ) is the stress operator defined by the formula:


|| ||
T(∂ x, n x ) = ||Ti j (∂ x, n(x))||,

where, in turn,
82 5 Mathematical Models of the Theory of Elasticity

∂ ∂ ∂
Ti j (∂ x, n(x)) = λn i (x) + μn j (x) + μδ i j .
∂x j ∂ xi ∂n(x)

In condition (5.29), the normal at point x is selected so that in the limit n(x) →
n(y), where n(y) is the external normal to surface S.
It should be noted that in the second boundary problem, the displacements are
determined by integral equations which are accurate to the displacement of the body
as a rigid. To eliminate this uncertainty, attention should be paid to the conditions
for the existence of the single solution. For example, when constructing a system of
resolution, one should take into account either the immobility of the vicinity of a
fixed-point M, or the immobility of the plane in the case of symmetrically deformable
bodies, etc.
Setting the boundary problem with conditions (5.24) refers to the third or mixed
boundary problem, that is, the mixed problem is characterized by setting the condi-
tions (5.28) on a part of the boundary surface, and the conditions (5.29) on the rest
of it.

Remark In the second and mixed problems of the theory of elasticity for the regions
with finite dimensions, the boundary conditions must be supplemented by the solv-
ability conditions of these problems, which consist in the requirement of the self-
equilibrium of the applied load. For a mixed task, unknown forces acting on Su must
be added to the applied load on Sσ :
∫ ∫
f (y)dS = 0; r × f (y)dS = 0.
S S

If there are several boundary surfaces, it is necessary to require the self-equilibrium


of the applied load as a whole rather than each surface.
Lamé equations are widely used in many methods for solving the problems of
theory of elasticity, since it does not need the equations of the strain compatibility,
which in this case are satisfied identity.
Particularly important problem is the homogeneous problem of the theory of
elasticity, when body forces can be zero.
We get some properties of displacements ui , resulting from (5.25), in the absence
of body forces, that is, F i = 0. We differentiate each of Eq. (5.25) by the corresponding
coordinate x i and perform convolution (summation by index i). We get:

(λ + μ)θ,ii +μΔ u i ,i = 0. (5.30)

Because:

θ,ii = θ,11 +θ,22 +θ,33 = Δ θ, Δ u i ,i = Δ (u 1 ,1 +u 2 ,2 +u 3 ,3 ) = Δ θ,

then from (5.28), we get


5.7 Boundary Problems of the Theory of Elasticity in Stresses 83

Δ θ = 0. (5.31)

Thus, the volume strain in an elastic isotropic solid with the absence of mass
forces is a harmonic function, which satisfies the Laplace’s Eq. (5.31).
We take now the Laplace operator from the left side of the Lamé equation,
assuming F i = 0:

(λ + μ)Δ θ,i +μΔ Δ u i = 0.

From which, taking into account (5.31), we get that Δ θ,i = (Δ θ),i = 0. Therefore:

Δ Δ u i = 0. (5.32)

Thus, each of the components of displacement vector is a biharmonic function of


the coordinates (satisfying the double Laplace’s equation).
However, it should not be thought, that the problem of the theory of elasticity can
be reduced to the integration of the system (5.32) or that the value θ can be found
using the known methods for solving the Laplace’s equation. To define a biharmonic
function, two conditions must be specified at the domain boundary (for example,
ui and ∂u i /∂n), while to solve the system (5.25), it is enough to set only the ui
values at each point on the surface. In addition, it is relatively easy to construct
three biharmonic functions with set values at the boundary, but they may not satisfy
Eq. (5.25). Finally, the system (5.32) is of the twelfth order, whereas the initial system
(5.25) is of the sixth order.
In addition, the value θ is never set at the boundary. It cannot be determined by
solving the Dirichlet6 problem.

Remark The system order can be defined as the product of the maximal derivative
order and the number of equations.

It should be noted that the described boundary problem for incompressible


materials (ν = 0.5) loses its ellipticity and its solution may not be unique.

5.7 Boundary Problems of the Theory of Elasticity


in Stresses

The theory of elastic problem can be set not only on displacements, but also on
stresses. This is useful when the body boundary is loading rather than displacements
and the stress state must be determined first.

6 Dirichlet Peter Gustav (1805–1859), German mathematician. The Dirichlet problem is the search
for a function satisfying the Laplace equation inside a domain and given at its boundary. In the
Neumann problem, the function derivative is given on the boundary.
84 5 Mathematical Models of the Theory of Elasticity

Stress problem as a boundary problem for an overridden system and its


simplification.
The resolving system of equations for the theoretical problems of elasticity in
stresses consists of three equilibrium equations in the form of Cauchy, six strain
compatibility equations and six ratios of the Hooke’s law. In total, these are 15
equations relative to 12 unknowns σi j , εi j , i, j = 1, 3 to which three boundary
conditions in the forces are attached.
Since this system obviously does not include displacements, it is natural to expect
that in this formula the problem in stresses has the unique solution. In fact, the
uniqueness theorem of the problem in stresses with this approach has been proved
by A.N. Konovalov.7
Using Hooke’s law, we exclude strains from the compatibility relations. As a
result, we obtain the following system to determine the unknown components of the
stress tensor:


3
∂ j σi j + f i = 0, i = 1, 3, (5.33)
j=1

ν ∑ 3
ν ∑ 3
Δ σi j + ∂i ∂ j σkk = − δi j ∂k f k
1+ν k=1
1 − ν k=1
− (∂i f j + ∂ j f i ), i, j=1, 3 (5.34)

The system of (5.33)–(5.34), as before, is a redefined system of equations (six


unknowns σi j = σ ji , i, j = 1, 3 satisfy nine equations: three equilibrium Eq. (5.33)
of the first order and six equations Beltrami8 -Michell9 (5.34) of the second order.
Equilibrium Eq. (5.33) was used in obtaining the Beltrami-Michell’s equations.
Therefore, Eq. (5.34) is always satisfied, if as σi j i, j = 1, 3, to take the system (5.33)
solutions. The opposite is not always true, since the differentiation of Eq. (5.33) used
in the construction of the Beltrami-Michell’s equations led to an expansion of the class
of functions σi j , i, j = 1, 3. Therefore, equilibrium Eq. (5.33) must be involved in
solving the theory of elasticity in stresses. This enables us to distinguish from the
Beltrami-Michell’s equations which class of solutions that lead to physically justified
stress–strain states of solids.
The definition of (5.33)–(5.34) leads to two important problems for the theory of
elasticity. The first problem is the study of the system (5.33)–(5.34) for compatibility.

7 Konovalov Anatoly Nikolaevich, Soviet and Russian scientist, Academician of the Russian
Academy of Sciences; one of the main areas of research is mathematical models and numerical
methods for problems of continua mechanics.
8 Eugenio Beltrami (1835–1900), Italian mathematician notable for his work concerning differential

geometry and mathematical physics.


9 John Henry Michell (1863–1940), Australian mathematician, Professor of Mathematics at the

University of Melbourne.
5.7 Boundary Problems of the Theory of Elasticity in Stresses 85

The second problem is the separation from the system (5.33)–(5.34) of the principal
system of six equations, which leads to the solution of the problem in stresses.
To separate a basic system of equations
Consider the problem of separation from the (5.33)–(5.34) basic system of equations.

Remark Detailed information on solving the problems of solid mechanics in stresses


is available in monographs.

First version. We select Eq. (5.34) as the basic system and satisfy Eq. (5.33) at
the boundary. We come to this formula of the problem in stress:
Determine the solution of such system in the area D occupied by the elastic body:

ν ∑ 3
ν
Δ σi j + ∂i ∂ j σkk = − δi j
1+ν k=1
1−ν

3
∂k f k − (∂i f j + ∂ j f i ); i, j = 1, 3,
k=1

when the following conditions are met at the boundary S of the elastic body:


3 ∑
3
n j σi j = Pni , ∂ j σi j = − f i ; i = 1, 3.
j=1 j=1

In this formula, the equilibrium equations of stress problems are used to release
a unique solution to the system (5.33)–(5.34).
Second version. As the basic system of equations for the stress problem, in addition
to the previously mentioned, a system consisting of three equilibrium equations and
three Beltrami-Michell’s equations with respect to the tangents of the stress tensor
components can be used:
⎧ 3

⎪ ∑

⎪ ∂ j σi j + f i = 0


j=1
(5.35)

⎪ ∑ 3

⎪ 1

⎩ Δ σi j + ∂i ∂ j σkk = −(∂ j f i + ∂i f i ); i /= j = 1, 3.
1+ν k=1

Therefore, for the problems of solid mechanics in stresses, the two formulas of
boundary problems are physically reliable.
86 5 Mathematical Models of the Theory of Elasticity

5.8 Homogeneous Problem of the Theory of Elasticity


in Stresses

The differential equilibrium equations at F i = 0 take the form:

σi j , j = 0. (5.36)

These three equations are not sufficient to determine the stress state in the elastic
body (six independent components of the stress tensor). As additional equations, we
can use the Beltrami-Michell’s equations in the absence of mass forces (compatibility
equations in stresses):

(1 + ν)Δ σi j + 3σ,i j = 0. (5.37)

Stresses, in the absence of mass forces, as well as components of the displacement


vector, have the properties (5.31) and (5.32). Indeed, summing the Eq. (5.37) at i =
j, we come to the equation:

Δ σ = 0. (5.38)

where σ = (σ11 + σ22 + σ33 )/3 is the average hydrostatic stress.


Taking into account (5.38), after applying the Laplace operator to the Eq. (5.37),
it follows that the stresses in static problems without mass forces are biharmonic
functions:

Δ Δ σi j = 0. (5.39)

5.9 The Problem of the Setting of Theory of Elasticity

A large class of the theory of elastic problems is convenient to solve not only in
Cartesian, but also in polar, cylindric or spherical coordinate systems. Let’s give
the main relations for the theory of elasticity in cylindrical and spherical coordinate
systems.
The cylindrical coordinates r, ϕ, and z are associated with Cartesian in the
following way, (Fig. 5.2):

x1 = x = r cos ϕ, x2 = y = r sin ϕ, x3 = z = z.

The linear element is given by the following formula:


5.9 The Problem of the Setting of Theory of Elasticity 87

Fig. 5.2 Relationship of


cylindrical and Cartesian z P(r, , z)
coordinates

r
z

y
x r

(ds)2 = (dr )2 + r 2 (dϕ)2 + (dz)2 .

Vector components of displacements:

u 1 = ur , u 2 = u ϕ , u 3 = u z .

Strain tensor components:

ε11 = εrr , ε22 = εϕϕ , ε33 = εzz ,


ε12 = εr ϕ , ε23 = εϕz , ε31 = εzr .

The following kinematic relations are valid:

∂u r 1 ∂u ϕ ur ∂u z
εrr = , εϕϕ = + , εzz = ,
∂r r ∂ϕ r ∂z
( ) ( ) ( )
1 1 ∂u r ∂u ϕ uϕ 1 ∂u ϕ 1 ∂u z 1 ∂u z ∂u r
εr ϕ = + − , εϕz = + , εr z = + .
2 r ∂ϕ ∂r r 2 ∂z r ∂ϕ 2 ∂r ∂z

The components of the stress tensor σrr , σϕϕ , σzz , τr ϕ , τϕz , τzr , are associated with
strains of the Hooke’s law with Lamé constants:

σrr = 2μεrr + λθ, . . . τr ϕ = 2μεr ϕ , . . . ,

where θ is the relative volumetric strain: θ = εrr + εϕϕ + εzz .


Equilibrium equations in stresses have the form:
88 5 Mathematical Models of the Theory of Elasticity

Fig. 5.3 Relationship of


spherical and Cartesian
z P(R, , )
coordinates

y
x r

1 ∂ 1 ∂τrϕ ∂τrz σϕϕ


(r σrr ) + + + ρFr = 0,
r ∂r r ∂ϕ ∂z r
1 ∂ 2 1 ∂σϕϕ ∂τϕz
(r τrϕ ) + + + ρFϕ = 0,
r ∂r
2 r ∂ϕ ∂z
1 ∂ 1 ∂τϕz ∂σzz
(r τr z ) + + + ρFz = 0,
r ∂r r ∂ϕ ∂z

where ρF r , ρF ϕ , ρF z are vector components of the body force.


In spherical coordinates, such conversion formulas are valid, (Fig. 5.3):

x1 = x = R sin ϑ cos ϕ, x2 = y = R sin ϑ sin ϕ, x3 = z = R cos ϑ.

Linear element: (ds)2 = (dR)2 + R 2 sin2 ϑ(dϕ)2 + R 2 (dϑ)2 .


Displacements: u 1 = u R , u 2 = u ϕ , u 3 = u ϑ .
Strain tensor components:

ε11 = ε R R , ε22 = εϕϕ , ε33 = εϑϑ ,


ε12 = ε Rϕ , ε23 = εϕϑ , ε31 = ε Rϑ .

The following kinematic relations are valid:

∂u R 1 ∂u ϕ uR uϕ 1 ∂u ϕ uR
εR R = , εϕϕ = + + ctgϑ , εϑϑ = + ,
∂r R sin ϑ ∂ϕ R R R ∂ϑ R
( ) ( )
1 1 ∂u R ∂u ϕ uϕ 1 1 ∂u R ∂u ϕ uϕ
ε Rϕ = + − , ε Rϑ = + − ,
2 R sin ϑ ∂ϕ ∂R R 2 R ∂ϑ ∂R R
( )
1 1 ∂u ϑ 1 ∂u ϕ uϕ
εϕϑ = + − ctgϑ .
2 R sin ϑ ∂ϕ R ∂ϑ R
5.10 2D-Problems of the Theory of Elasticity 89

The components of the stress tensor σ R R , σϕϕ , σϑϑ , τ Rϕ , τϕϑ , τϑR are associated
with strains of the Hooke’s law with Lamé constants:

σ R R = 2με R R + λθ, . . . , τ Rϕ = 2με Rϕ , . . . ,

where θ = ε R R + εϕϕ + εϑϑ .


Equilibrium equations:

1 ∂(R 2 σ R R ) 1 ∂(τ Rϕ sin ϕ) 1 ∂τ Rϑ σϕϕ + σϑϑ


+ + − + ρFR = 0,
R ∂R R sin ϕ ∂ϕ R sin ϕ ∂ϑ R
1 ∂(R 3 τ Rϕ ) 1 ∂(τϕϑ sin2 ϑ) 1 ∂σϕϕ
+ + + ρFϕ = 0,
R 3 ∂R R sin ϕ
2 ∂ϕ R sin ϕ ∂ϕ
1 ∂(R 3 τ Rϑ ) 1 ∂(σϑϑ sin ϑ) 1 ∂τϕϑ ctgϑ
+ + − σϕϕ + ρFϑ = 0.
R 3 ∂R R sin ϕ ∂ϑ R sin ϑ ∂ϕ R

5.10 2D-Problems of the Theory of Elasticity

The problem that the stress–strain state functions only depend on two coordinates is
very important for applications. These are 2D-problems of the theory of elasticity. At
the same time, two cases are distinguished: plane strain state and plane stress state.
• Plane strain state. This case corresponds to the stress–strain state formation in a
long prismatic body (with a longitudinal axis of coordinate z) loaded with surface
forces independent of z and has no component along this axis.
As an example, the cylinder loaded by forces p1 , p2 , p3 , which linearly distributed
along its general axis is shown in the Fig. 5.4. The elastic body may be either infinitely
long or of finite length, but its edges shall be appropriately fixed. In this case, the
stress–strain state on the cross section of each body is plane strain state.
In the future, simplified relationships will be used, so the tensor form of recording
will only be used for compact writing of formulas. Displacements in Cartesian
coordinate system x, y, z are defined through u, v, w, respectively:

u = u(x, y), v = v(x, y), w = const or w = 0. (5.40)

In addition, all the derivatives of displacements by z (by x 3 ) are zeros. The


deformas, including the volume deformation, will be as follows:
( )
∂u ∂v 1 ∂u ∂v
εx x = , ε yy = , εx y = + ,
∂x ∂y 2 ∂y ∂x
90 5 Mathematical Models of the Theory of Elasticity

Fig. 5.4 Example of plane p1


strain state

y
p3
p
2

z x

∂u ∂v
εx z = ε yz = εzz = 0, θ = + . (5.41)
∂x ∂y

Only one strain compatibility condition remains:

∂ 2 εx x ∂ 2 ε yy ∂ 2 εx y
+ =2 . (5.42)
∂y 2 ∂x 2 ∂ x∂ y

Non-zero stress tensor components are σx x , σx y , σ yy , σzz .


The formulas of the generalized Hooke’s law take the form:

σx x = 2με x x + λθ , σ yy = 2με yy + λθ , σx y = 2με x y , σzz = λθ . (5.43)

Strains are expressed through stresses by the following formulas [obtained from
(5.6)]:

1[ ] 1[ ]
εx x = σx x − ν(σ yy + σzz ) , ε yy = σ yy − ν(σzz + σx x ) ,
E E
1+ν
εx y = σx y .
E
It is easy to show that for a plane strain of a body, the number of independent
components of the stress tensor is three. In fact, from the Hooke’s law, under the
condition εzz = 0, it is as follows:

σzz = ν(σx x + σ yy ). (5.44)

Only the following equations remain from equilibrium equations:

∂σx x ∂σx y ∂σ yy ∂σx y


+ = 0, + = 0. (5.45)
∂x ∂y ∂y ∂x
5.10 2D-Problems of the Theory of Elasticity 91

body forces are not taken into account here.


The boundary conditions for stresses, if Rx and Ry are set at the cylinder edge,
are:

σx x cos(ν, x) + σx y cos(ν, y) = Rx , σx y cos(ν, x) + σ yy cos(ν, y) = R y . (5.46)

For a prismatic body with a length of l, the end sections of which are supported,
the boundary conditions have the form:

w(x, y, 0) = w(x, y, l) = 0.

Then the forces acting on the cross sections of the end equal to:

Pz = σzz dS,
S

where stresses are integrated along the cross section of the prismatic body.
Equilibrium equations in displacements (Lamé’s equations) (5.25) give the form:
( ) ( )
1 ∂ ∂u ∂v 1 ∂ ∂u ∂v
Δ u + + = 0, Δ v + + = 0.
1 − 2ν ∂ x ∂x ∂y 1 − 2ν ∂ y ∂ x ∂y

Accordingly, the compatibility equations in stresses (Beltrami-Michell’s equa-


tions) (5.37) become as follows:

3 ∂ 2σ
Δ σx x + = 0,
1 + ν ∂x2
3 ∂ 2σ
Δ σ yy + = 0,
1 + ν ∂ y2
3 ∂ 2σ
Δ σx y + = 0,
1 + ν ∂ x∂ y

where the average stress is (the expression (5.27) for σz is taken into account):

1 1+ν
σ= (σx x + σ yy + σzz ) = (σx x + σ yy ).
3 3
If the prismatic body is finite in length and there is no load in its ends, then
the conditions of plane strain are broken. However, even in this case, for sections
sufficiently remote from the ends, on the basis of the Saint-Venan principle, the
stress–strain state can be considered corresponding to the plane strain conditions.
• Plane stress state. In this case, we are talking about a flat elastic body of
small thickness (plate), which is only loaded by forces on its plane, and has
no normal stresses in the thickness direction, (Fig. 5.5). The applied forces are
92 5 Mathematical Models of the Theory of Elasticity

Fig. 5.5 Example of plane


P3 z
stress state

P1 P2
h

evenly distributed across the thickness and are independent of z (which is always
done with good approximation for thin plates). Forces can also be distributed
symmetrically with respect to the median plane (an imaginary surface that bisects
the thickness) of the plate. In this case, their values of average plate thickness can
be entered.
With a plane stress state in the (x, y) plane, there are the following non-zero stress
tensor components:

σx x = σx x (x, y), σ yy = σ yy (x, y), σx y = σx y (x, y).

The remaining components are zeros: σzz = σx z = σ yz = 0.


The components of displacements u, v, and w in this case are independent of the
z coordinate.
In the case of a plane stress state, kinematic equations correspond to Eq. (5.41) for
a plane strain state, compatibility conditions correspond to Eq. (5.42), equilibrium
equations correspond to Eq. (5.45), and boundary conditions correspond to conditions
(5.46).
When considering strains, the difference between plane strain state and plane
stress state is exerted, for example, in Hooke’s law. Since σzz = 0, it follows from
Hooke’s law:
1[ ] 1[ ]
εx x = σx x − νσ yy , ε yy = σ yy − νσ x x ,
E E
ν σx y
εzz = − (σx x + σ yy ), εx y = . (5.47)
E 2G
Reversing (5.47), yield stress expressions through strains are:

E
σx x = (εx x + νε yy ),
1 − ν2
5.10 2D-Problems of the Theory of Elasticity 93

E E
σ yy = (ε yy + νε x x ), σx y = εx y .
1−ν 2 1+ν

These formulas differ from the corresponding formulas (5.43) and (5.44).
As already mentioned, the plane stress state is realized not precisely, but only
approximately in thin plates.
Filon10 showed that it is possible to consider a generalized case and modify the
given ratios. Symmetric distribution of applied forces relative to median plane of
plate is assumed. If it is assumed that the stresses σzz in the plate are absent, and
on the outer surfaces of the plate, σzx and σzy are zeros, then for displacements and
stresses, their average values are calculated for the thickness of the plate, for example:

∫ h
1
σ∗x x = σx x dz,
2h
−h

where 2h is the plate thickness.


As a result, the task is reduced to the previous task. Such a stress state is called a
generalized plane stress state.
Stress compatibility equations for plane stress problems
To solve the plane problem in stresses, strain compatibility Eq. (5.42) must be written
in stresses, using the Hooke’s law (5.47). As a result of this operation, we obtain:

1 ∂2 [ ] 1 ∂2 [ ] 1 ∂ 2 σx y
σ x x − νσ yy + σ yy − νσ x x − 2 = 0.
E ∂ y2 E ∂x2 2G ∂ x∂ y

Next, we express the derivatives of tangent stresses using equilibrium Eq. (5.45):

∂ 2 σx y ∂ 2 σx x ∂ 2 σx y ∂ 2 σ yy
=− , =−
∂ x∂ y ∂x2 ∂ x∂ y ∂ y2

and substitute these ratios into the previous ratio. After replacing the shear modulus
G with the expression 2G = E/(1 + v), we have:
( 2 )
∂2 [ ] ∂2 [ ] ∂ σx x ∂ 2 σ yy
σ xx − νσ yy + σ yy − νσ xx + (1 + ν) + = 0.
∂ y2 ∂x2 ∂x2 ∂ y2

After reductions:
( )
∂2 ∂2
+ 2 (σx x + σ yy ) = 0,
∂x2 ∂y

10 Filon L. N. G., 1875–1937, English mathematician, physicist.


94 5 Mathematical Models of the Theory of Elasticity

or, using the Laplace operator in a planar coordinate system Δ ≡ ∂ 2 /∂ x 2 + ∂ 2 /∂ y 2 ,


we get:

Δ (σx x + σ yy ) = 0. (5.48)

Therefore, the sum of stresses is a harmonic function.


• Airy11 stress function
In the absence of mass forces, equilibrium Eq. (5.45) are usually satisfied by the
introduction of a stress function in accordance with the formulas:

∂ 2ϕ ∂ 2ϕ ∂ 2ϕ
σx x = , σ yy = , σ xy = − . (5.49)
∂ y2 ∂x2 ∂ x∂ y

The function ϕ(x, y) is called the Airy function.


Sum up normal stresses:

∂ 2ϕ ∂ 2ϕ
σx x + σ yy = + = Δ ϕ.
∂x2 ∂ y2

Take the Laplace operator from both parts of the obtained equality. Since stress
compatibility equations must be satisfied (5.48), this leads to a main differential
equation to determine the Airy function:

∂ 4ϕ ∂ 4ϕ ∂ 4ϕ
Δ Δ ϕ = + 2 + = 0. (5.50)
∂x4 ∂ x 2∂ y2 ∂ y4

Equation (5.50) is a biharmonic equation and its solution is biharmonic functions.


Many of its partial solutions are known, each of which corresponds to a certain stress
state satisfying equilibrium equation and compatibility equation. For example:

x 2 , y 2 , x y, x 3 , y 3 , x 2 y, x y 2 , x 4 − y 4 ,
cos(λx)ch(λy), cos(λy)ch(λx), . . .

The main difficulty of the process of solution building is the selection of functions
that satisfy boundary conditions. Numerous problems of the theory of elasticity,
which are of great practical importance, have been solved taking into account specific
boundary conditions. However, there are no general solutions to the biharmonic
equation, and there are also no general methods for solving it. In order to close the
mathematical boundary problem by defining the Airy function, boundary conditions
and displacements must be expressed through it.

11 Sir George Biddell Airy (1801–1892) was an English mathematician and astronomer. His many
achievements include work on planetary orbits, a method of solution of two-dimensional problems
in solid mechanics.
5.10 2D-Problems of the Theory of Elasticity 95

Thus, in order to obtain an exact solution to the theory of elastic problem, it is


necessary to find such functions that, in addition to satisfy the biharmonic Eq. (5.50),
would also strictly satisfy the boundary conditions at each point of the solid surface.
In detail, the plane problem of the theory of elasticity is considered in many
textbooks. It should be noted that the classical work of N.I. Muskhelishvili.
According to the theory of elasticity, much literature has been published.

Control Questions

1. Hooke’s generalized law for an isotropic body:


1) σi j = 2μεi j + λθδi j ;
2) U = σi j εi j /2;
3) εαα,ββ + εββ,αα − 2εαβ,αβ = 0.
2. Hooke’s generalized law for an anisotropic body:
1) U = σi j εi j /2;
2) σi j = E i jkl εkl ;
3) σi j = 2μεi j + λθδi j .
3. Clapeyron Formula:
[ ]
1) εi j = 2G
1
σi j − 1+ν

σδ i j ;
2) σi j = E i jkl εkl ;
3) U = σi j εi j /2.
4. Total potential strain energy U:
( )
1) U = 1+ν6E
(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 ;
2) U = 1−2ν6E( (σ1
+ σ 2 + σ 3 )2 ;
)
3) U = 2E 1
σ12 + σ22 + σ32 − 2ν(σ1 σ2 + σ2 σ3 + σ3 σ1 ) .
5. Temperature is included in expressions linking stresses and strains:
1) deviators;
2) spherical parts;
3) intensities.
6. In the absence of mass forces, displacements are:
1) harmonic functions;
2) biharmonic functions;
3) generalize functions.
7. Potential Energy of shape change Uf :
( )
1) U f = 1+ν
6E
(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 ;
2) U f = 1−2ν
6E( (σ1
+ σ 2 + σ 3 )2 ;
)
3) U f = 2E σ1 + σ22 + σ32 − 2ν(σ1 σ2 + σ2 σ3 + σ3 σ1 ) .
1 2
96 5 Mathematical Models of the Theory of Elasticity

8. Potential energy of volume change UV :


( )
1) UV = 1+ν 6E
(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 ;
2) UV = 1−2ν6E( (σ1
+ σ2 + σ3 )2 ;
)
3) UV = 2E σ1 + σ22 + σ32 − 2ν(σ1 σ2 + σ2 σ3 + σ3 σ1 ) .
1 2

9. Lamé’s equations are equilibrium equations in:


1) stresses;
2) strains;
3) displacements.
10. Beltrami-Michell equations are equations:
1) equilibrium;
2) compatibility;
3) expressions for elastic potential.
11. Lamé’s equations:
[ ]
1) εi j = 2G
1
σi j − 1+ν

σδ i j ;
2) σi j = E i jkl εkl ;
3) (λ + μ)θ,i +μΔ u i + ρFi = 0
12. Beltrami-Michell equations:
[ ]
1) εi j = 2G 1
σi j − 1+ν

σδi j ;
2) σi j = E i jkl εkl ;
ν Σ ν Σ
3) Δ σi j + 1+ν ∂i ∂ j 3k=1 σkk = − 1−ν δi j 3k=1 ∂k f k − (∂i f j + ∂ j f i ), i, j =
1, 3
13. In the absence of mass forces, the displacements satisfy the equations:
1) harmonic equations;
2) biharmonic equations;
3) Bessel equations.
14. Boundary conditions are specified on the surface:
1) deformed;
2) not deformed;
3) on both surfaces.
15. On the surface of the body carry out:
1) equilibrium equations;
2) initial conditions;
3) boundary conditions.
16. What are the boundary problems of the theory of elasticity in displacements?
What are the first, second and mixed problems of the theory of elasticity in
displacements?
5.10 2D-Problems of the Theory of Elasticity 97

17. What are the boundary problems of the theory of elasticity in stresses? What is
the definition of an overridden system?
18. Stresses σzz = 0 in case of plane state:
1) deformed;
2) loading;
3) in both cases.
19. In case of plane strain state:
1) u = constant, v = constant, w = constant;
2) u = (x, y),v = (x, y),w = constant;
3) u = u(x, y, z), v = v(x, y, z), w = (x, y, z).
20. In case of plane strain state:
1) εzz = εx z = ε yz = 0;
2) εzz = constant, εx z = constant, ε yz = constant;
3) εzz = εzz (x, y), εx z = εx z (x, y), ε yz = ε yz (x, y).
21. Compatibility equations at plane stress state:
( )
1) εx x = ∂∂ux , ε yy = ∂v
∂y
, ε xy = 1 ∂u
2 ∂y
+ ∂v
∂x
;
∂u ∂v
2) θ = ∂x
+ ∂y
;
∂ 2 εx x ∂ 2 ε yy ∂2ε
3) ∂ y2
+ ∂x2
= 2 ∂ x ∂x yy .
22. For plane strain state, the number of independent stress components:
1) three;
2) four;
3) six.
23. Equilibrium equations at plane strain state:
∂ 2 εx x ∂ 2 ε yy ∂2ε
1) ∂ y2
+
∂x2
= 2 ∂ x ∂x yy ;
∂σx x ∂σ ∂σ ∂σx y
2) ∂x
+ ∂ yx y = 0, ∂ yyy + ∂x
= 0,
( )
3) εx x = ∂∂ux , ε yy = ∂v ∂y
, εx y = 1
2
∂u
∂y
+ ∂v
∂x
.

24. For plane stress state:


1) σx x = constant, σ yy = constant, σzz = constant;
2) σzz = σx z = σ yz = 0;
3) σx x = σx x (x, y, z), σ yy = σ yy (x, y, z), σx y = σx y (x, y, z).
25. Airy stress function is defined:
∂σx x ∂σx y ∂σ ∂σ
1) ∂x
+ ∂y
= 0, ∂ yyy + ∂ xx y = 0;
( )
∂u
2) εx x = ∂x
, ε yy = ∂v
∂y
, εx y = 21 ∂u∂y
+ ∂∂vx ;
∂2ϕ
, σ yy = ∂∂ xϕ2 , σx y = − ∂∂x∂ϕy .
2 2
3) σx x = ∂ y2
98 5 Mathematical Models of the Theory of Elasticity

26. Airy stress function satisfies the equation:


1) harmonic;
2) biharmonic;
3) Laplace’s.

Control Exercises

1. Show that for isotropic elastic media:


ν λ
1
a) 1+ν = 2(λ+μ)
3λ+2μ
; b) 1−ν = λ+2μ 2μν
; c) 1−2ν = 1+ν
3kν
; d) 2(1 + ν)μ =
3k(1 − 2ν).
2. Let the axis x 1 be the symmetry axis of order N = 2. Define the type of elastic
constant matrix Cαβ , considering that Cαβ = Cβα .
Answer:
⎡ ⎤
C11 C12 C13 C14 0 0
⎢C C C C ⎥
⎢ 12 22 23 24 0 0 ⎥
[ ] ⎢ ⎢C C C C 0 0 ⎥

Cαβ = ⎢ 13 23 33 34 ⎥
⎢ C14 C24 C34 C44 0 0 ⎥
⎢ ⎥
⎣ 0 0 0 0 C55 C56 ⎦
0 0 0 0 C56 C66

3. Let the axes that make up equal angles with coordinate axes be axes of elastic
symmetry of order N = 3. Show that there are twelve independent elastic
constants and the matrix of elastic constants has the form:
⎡ ⎤
C11 C12 C13 C14 C15 C16
⎢C C C C C C ⎥
⎢ 13 11 12 16 14 15 ⎥
[ ] ⎢ ⎢C C C C C C ⎥

Cαβ = ⎢ 12 13 11 15 16 14 ⎥.
⎢ C41 C42 C43 C44 C45 C46 ⎥
⎢ ⎥
⎣ C43 C41 C42 C46 C44 C45 ⎦
C42 C43 C41 C45 C46 C44

4. For an elastic body, let axis x 3 be the axis of elastic symmetry of order N = 6.
Show that only the following elastic constants are non-zero:

C11 = C22 , C33 , C55 = C44 , C66 = 0, 5(C11 − C12 ), C13 = C23 .

5. Let the stress field for the medium be:


⎡ ⎤
[ ] x1 + x2 σ12 0
σi j = ⎣ σ12 x1 − x2 0 ⎦,
0 0 x2

where σ12 is function from x 1 and x 2 .


5.10 2D-Problems of the Theory of Elasticity 99

Let equilibrium equations be satisfied in the(absence


)
Λ
of mass forces and let
e1 Λ Λ

the stress vector in the plane x1 = 1 be given as t = (1 + x2 )e1 +(6 − x2 )e2 .


Define σ12 as a function from x 1 and x 2 .
Answer: σ12 = x1 − x2 + 5.
6. The displacement field is given through some vectors qi by the following
equation:
( )
u i = 2(1 − ν)qi, j j − q j, ji /G.

Show that the Lamé’s equations are satisfied if the mass forces are zeros and
qi are biharmonic functions such that qi, j jkk = 0.
Let q1 = x2 /r and q2 = −x1 /r (r 2 = xi xi ). Define the resulting stress field.
Answer:
( )
σ11 = −σ22 = 6Qx1 x2 /r 5 ; σ33 = 0; σ12 = 3QG x22 − x12 /r 5 ;
σ13 = −σ31 = 3Qx2 x3 /r 5 ; σ23 = 0, Q = 4(1 − ν)/G.

7. Show what ϕ(x1 , x2 ) = x14 x2 +4x12 x23 −x25 is the Airy stress function. Calculate
the stress tensor for this case under the assumption of a plane strain state with
Poisson’s ratio ν = 0.25.
Answer:
⎡ ⎤
[ ] 24x12 x2 − 20x23 −4x13 − 24x1 x22 0
σi j = ⎣ −4x13 − 24x1 x22 12x12 x2 + 8x23 0 ⎦.
0 0 9x1 x2 − 3x2
2 3

8. Airy stress function has the form:

ϕ(x1 , x2 ) = Dx12 x23 + F x25 .

Show that in this case F = −D/5.


9. ) has the form: ϕ = ϕ2 + ϕ3 + ϕ5 , where ϕ(x1 , x2 ) =
The( stress function
D5 x12 x23 − 0.2x25 + 0.5B3 x12 x2 + 0.5A2 x12 x2 .
Show that for this stress function the stress components are defined as:
( )
σ11 = D5 6x12 x2 − 4x23 ; σ22 = 2D5 x23 + B3 x2 + A2 ; σ12 = −6D5 x1 x22 − B3 x1 .

10. Beam with rectangular cross section of unit width and length 2L loaded with
distributed load of intensity q (Fig. 5.6). Transverse forces act on both ends of
beam V.
List the six boundary conditions for this beam that the stresses must satisfy.
Answer:
1) σ22 = −q, x2 = +c; 2) σ22 = 0, x2 = −c; 3) σ12 = 0, x2 = ±c;
100 5 Mathematical Models of the Theory of Elasticity

Fig. 5.6 Illustration for exercise 10

∫ c ∫ c
4) −c σ12 dx2 = q L , x1 = ±L; 5) −c σ11 dx2 = 0, x1 = ±L; 6)
∫ c
−c σ11 x 2 dx 2 = q L , x 1 = ±L.
11. Using the boundary conditions 1, 2, and 3 listed in exercise 10, show that for
stress expressions in exercise 9, you need to:
( )
A2 = −q/2; B3 = −3q/(4c); D5 = q/ 8c3 .

In this case, the formulas for stresses have the form:


( ) ( )
q 2 3 q 1 3 2 3
σ11 = x x2 − x2 ; σ22 =
2
x − c x2 − c ;
2
2I 1 3 2I 3 2 3
q ( 2 )
σ12 = − x 1 x 2 + c2 x 1 ,
2I

where I = (2/3)c3 is plane moment of inertia of beam cross section.


12. Show that using the stresses calculated in exercise 11, the boundary conditions
4 and 5 of exercise 10 are satisfied and the boundary condition 6 is not satisfied.
13. Let’s continue exercises 11 and 12. We add an additional term to the stress
function for implementation of boundary condition 6:

ϕ = D3 x23 .

Show that from the boundary condition 6 it follows that:


( )
3q 1 L2
D3 = − 2 .
4c 15 6c

Therefore, finally:
( )
q 2 2 1 2 1 3
σ11 = x − x + c − L x2 .
2
2I 1 3 2 15 6
Chapter 6
Mathematical Models of Solid
with Rheological Properties

As mentioned repeatedly before, in order to construct a full system of equations for


describing the solid behavior, it is necessary to attach the special relations describing
the solid mechanical behavior (the relations between the solid stress–strain state) to
the constitutive equations.
In this section, we will talk about theories describing deformation processes in
solids with rheological properties.

6.1 Creep and Relaxation

In the theory of elasticity, it is assumed that external influences do not change over
time, so the solid stress–strain state remains unchanged over time. However, many
materials deform over time even under constant stresses. This property of materials
is called creep. Creep can result in significant changes in the stress–strain state of
solids over time.
The mechanical properties of materials that are significantly time-dependent are
called rheonomic properties.
Creep can be experimentally examined by the stretching of cylindrical samples.
Strain growth graphs over time at constant stresses are called creep curves. Test
conditions and creep curve are shown schematically in Fig. 6.1. The upper end of the
sample is fixed and the load is applied to the lower end. The strain change curve from
time (creep curve) is constructed on the results of the length change of the sample.
The strain increases from its initial value ε0 , which reflects the elastic properties of
the material and corresponds to the “instantaneous” applied load P, that is, ε0 is the
elastic initial strain.
We can distinguish three characteristic sections on the creep curve. In the first
section (AB), the creep speed is high, but decreases monotonically, starting from a
certain value. This is a section of unsteady creep. The second curve section (BC)
is almost straight, where the speed of creep is almost constant. This is a section of

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 101
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_6
102 6 Mathematical Models of Solid with Rheological Properties

Fig. 6.1 Creep curve

steady creep. The third section (CD) predates the destruction, Where the creep rate
increases monotonically.
It should be noted that creep is observed at any stresses. Therefore, it is impossible
to specify any ultimate load at the beginning of creep.
The phenomenon of reducing stresses in the body at a constant strain is called
relaxation. Suppose that the sample was “instantaneously” stretched so that its “use-
ful” length l 0 equal to l, (Fig. 6.2). To do this, the load P1 was required. Then we
fixed the sample in a stretched state for a period of time. After that, the sample was
released from fixation and a load was applied again, under the influence of which the
sample “useful” part was stretched to the value of l. In this case, we need a load P2
less than the original load P1 . Therefore, the stress required to maintain a constant
strain is reduced (relaxed).
In solid deformable bodies, creep and relaxation, as rheonomic properties of
the material, take place simultaneously and are interconnected. The relationship
between creep and relaxation can be expressed analytically, introducing time into
the relationship of solid stresses and strains.
Viscoelasticity (relaxing) are solids with elastic and rheonomic properties. At the
same time, a viscous element prevails in such bodies, and strains grow indefinitely

Fig. 6.2 Relaxation curve


6.2 Solid Mathematical Models 103

under constant stresses. With constant strains, the initial stresses will decrease over
time (relax) in such bodies.
If the relationship between strain increments and stress increments at any moment
is almost linear, then this medium behavior can be described using the linear creep
model. In this case, the theory of linear hereditary media can be used to describe
the medium deformation over time. The complete strain at any moment consists of
two components: elastic strain at the loading and creep strain itself. Mathematically,
this can be expressed in the following form:

∫t
1 1
ε(t) = σ(t) + L(t, τ)σ(τ)dτ,
E E
0

where L(t, τ) is a function expressing hereditary material properties.

6.2 Solid Mathematical Models

6.2.1 Structural Rheological Models

In the most general form, the differential equation linking stress and strain can be
represented in the following form, while taking into account the development of
deformation processes over time:

dσ d2 σ dn σ
a0 σ + a1 + a2 2 + · · · + an n
dt dt dt
dε d2 ε dm ε
= b0 ε + b1 + b2 2 + · · · + bm m , (6.1)
dt dt dt
where ai (i = 0, 1, . . . , n), b j ( j = 0, 1, . . . , m) are factors determined by material
properties.
The most common approach to obtain a rheological equation for a specific mate-
rial today is to summarize a large number of experimental research data, construct
the corresponding creep curves, and compare them with the curves with the same
name based on theoretical models.
In practice, a common approach is to use only a few terms in both parts of Eq. (6.1)
to describe experimental data. This is equivalent to describe the viscoelastic behavior
of materials based on rheological structural models consisting of elastic elements
(their behavior is described using Hooke’s law) and viscous dampers (deformed
according to Newton’s law), (Fig. 6.3).
Structural units with elastic properties can be presented to springs with the Hooke’s
deformation law:
104 6 Mathematical Models of Solid with Rheological Properties

b
a

Fig. 6.3 Structural models of solids viscoelastic behavior: a viscoelastic (relaxing) model; b linear
elastic model with lag; c linear solid model; d model with elastic lag and stress relaxation

σ = E ε, (6.2)

where σ is the stress; ε is the strain; E is the modulus of elasticity (coefficient


proportionality).
A structural element that describes the material viscous behavior is called a
Newton element. It is a piston with through holes in a cylinder containing viscous
liquid (damper), (Fig. 6.3). This element corresponds to the simplest model of
material linear viscous behaviour—Newton’s law of linear viscous deformation:

∫t
dε σ(τ)
σ=η· or ε(t) = ε(0) + dτ, (6.3)
dt η
0

where dε/dt is the strain rate; η is the viscosity coefficient.


Material has creep at a constant speed under a constant value of loading.
6.2 Solid Mathematical Models 105

Fig. 6.4 Parallel connection


of elements

It should be noted that the Newton element cannot reproduce instantaneous elas-
ticity. In addition, the concept of initial length cannot be defined for the Newton
element, and it can accumulate unlimited deformation under constant loads.

Remark In “pure form” the Newton element corresponds to the liquid behavior.

Other linear viscoelastic models can be assembled as a parallel or serial connection


of springs (Hooke’s model) and pistons (dampers) (Newton’s model):
• When the elements are connected in parallel (Fig. 6.4), the total stress equals to
the sum of the stresses in the elements, and the strains in the elements are the
same and equal to the total structural strain:

σ = σ1 + σ2 , ε = ε1 = ε2 . (6.4)

• When connecting elements are consecutive (Fig. 6.5), the general rules are as
follows:

σ = σ1 = σ2 , ε = ε1 + ε2 . (6.5)

Arbitrary rheological models can be compiled according to these rules. Note that
there may be situations where different models describe the same material behavior
(with selected appropriate material properties).
Consider some “classic models”. Note that in this section, the models are described
in the one-dimensional case.
• Maxwell model
Maxwell model is one of the simplest models to describe the viscoelastic behavior
of solids. It represents a consecutive connection of a spring and the damper (Fig. 6.3a).
In this case, if an external force is applied at the initial moment, then the spring will
be instantly stretched (instantaneous elastic strains). If in the future the total stretch
of the system remains constant (constant strains), then the spring will shrink and

Fig. 6.5 Consecutive (serial) connection of elements


106 6 Mathematical Models of Solid with Rheological Properties

gradually pull up the piston. At the same time, the external force should obviously
decrease. Thus, Maxwell model illustrates the stress relaxation phenomenon.
In a one-dimensional Maxwell model, following the rule (6.5), the stresses in the
elements are the same, and the strains are summed up. If the equation of Hooke’s
law is time differentiated, we get the connection between stress and strain, and we
add a viscous component to the obtained strain rate:

σ̇ E ση
ε = ε E + εη ⇒ ε̇ = ε̇ E + ε̇η ⇒ ε̇ = + .
E η

Since σ = σ E = ση , we finally get:

σ̇ σ
ε̇ = + . (6.6)
E η

Let’s consider the process of limited stress action and their corresponding strains in
the period of time from 0 to t. Integrate the representation (6.6) and get the expression
for deformations:

∫t
1 1
ε(t) = ε(0) + [σ(t) − σ(0)] + σ(τ)dτ. (6.7)
E η
0

We get expressions for stresses by multiplying (6.6) by E exp((E/η)t) [Bertram,


A. Solid Mechanics. Theory, Modeling, and Problems/Albrecht Bertram, Rainer
Glüge—Springer International Publishing Switzerland, 2015. 331 p.]. As a result,
we obtain an expression that is a complete time derivative:
( ) ( )
E E
E ε̇(t) exp t = σ̇(t) exp t
η η
( ) ( ( ))
E E d E
+ σ(t) exp t = σ(t) exp t .
η η dt η

Integrate this expression over time between 0 and t:

∫t ( ) ( ( ))|t
E E |
E ε̇(τ) exp τ dτ = σ(τ) exp τ ||
η η τ=0
0
( )
E
= σ(t) exp t − σ(0).
η

Let’s rewrite this formula with respect to stresses:


6.2 Solid Mathematical Models 107

( ) ∫t ( )
E E
σ(t) = σ(0) exp − t + E ε̇(τ) exp − (τ − t) dτ. (6.8)
η η
0

The integral in the expression (6.9) can be rewritten as follows:

∫t ( ) ( )|t
E E ||
ε̇(τ) exp τ dτ = ε(τ) exp τ
η η |τ=0
0
∫t ( )
E E
− ε(τ) exp τ dτ.
η η
0

As a result, we have:
( )
E
σ(t) = σ(0) exp − t
η
( )[ ( ) ]
E E
+ E exp − t ε(t) exp t − ε(0)
η μ
∫t ( )
E2 E
− ε(τ) exp (τ − t) dτ,
μ η
0

or:
( )
E
σ(t) = Eε(t) + [σ(0) − E ε(0)] exp − t
η
∫ t ( )
E2 E
− ε(τ) exp (τ − t) dτ. (6.9)
μ η
0

The first term in (6.9) is the elastic component of the stresses. The second term
expresses the influence of initial values, which decreases over time according to the
exponential function. The third term is the convolution integral, which shows the
contribution of strains: more remote events have less impact at the current time than
more recent events (fading memory).
Material behavior within Maxwell model demonstrates the results of creep test
(Fig. 6.6, Bertram, A. Solid Mechanics. Theory, Modeling, and Problems/Albrecht
Bertram, Rainer Glüge—Springer International Publishing Switzerland, 2015. 331
p.).
It is assumed that before t = 0 the stresses are zeros (σ = 0) and after that they
are constant, i.e., σ = σk = constant, t > 0. We obtain it on the basis of (6.8), if
ε(0) = 0:
108 6 Mathematical Models of Solid with Rheological Properties

Fig. 6.6 Tests for creep (a) and relaxation (b) within the Maxwell model

σk σk
ε= + t, t > 0. (6.10)
E η

From (6.11), it follows that there is an unlimited linear creep. In addition, it


follows from (6.11) that strains from fast loads in this case are “extinguished” by
elastic strains (instantaneous elasticity).
In case there are constant strains εr (relaxation test), the damper will instantly
operate and the spring elastically stretches so that the initial stresses equal to σ0 =
Eεr . After that, according to (6.9), we have:
( )
E
σ(t) = Eεr exp − t .
η

Thus, the stresses tend to zero until the spring is unloaded due to the piston
displacements.
The material constant tr = η/E (dimension “time”) is called the relaxation time.
For t = tr , we get:

σ(tr ) = Eεr e−1 ≈ 0.368σ0 .

Using (6.6) (in this case ε̇ = 0) for t ≥ 0, we have:

σ(t) = −tr σ̇(t).


.
For t = 0, the initial value σ̇(t) is: σ(0) = −σ0 /tr .
Therefore, the relaxation time is the inverse of the relaxation rate.
The description of tensor of the Maxwell model can generally be presented as
follows:
6.2 Solid Mathematical Models 109
[ ( ) ]
∂σi j 1 ∂ei j 1 2 ∂θ
+ σi j = 2μ + δi j λ+ μ θ+λ .
∂t η ∂t η 3 ∂t

• Kelvin-Foigt model

Kelvin-Foigt model is a parallel connection between spring and damper. This


model clearly explains the phenomenon of creep (Fig. 6.3b). If a constant external
load (for example, tension) is applied to this model, then there will be no initial
instantaneous strains, and the viscous piston will move as long as the spring allows.
The Kelvin-Foigt body is an example of the simplest linear elastic-lag medium.
In the Kelvin Foigt model, according to (6.4), the stresses in the elements are
summed, and the strains in the elements are the same, i.e.,

ε = ε E = εη , σ E = Eε, σ = σ E + ση ,
. .
ση = η ε, ⇒ σ = Eε + η ε . (6.11)

To obtain the expression of strains, we apply to (6.11), the procedure we used to


obtain expressions for stresses in the case of Maxwell model.
We multiply (6.11) by η1 exp((E/η)t) [Bertram, A. Solid Mechanics. Theory,
Modeling, and Problems/Albrecht Bertram, Rainer Glüge—Springer International
Publishing Switzerland, 2015. 331 p.]. As a result, we get:
( ) ( ) ( ) ( ( ))
1 E E E . E d E
σ(t) exp t = ε(t) exp t + ε(t) exp t = ε(t) exp t .
η η η η η dt η

Integrating over time in the range from 0 to t, we have:

∫t ( ) ( )|t ( )
1 E E || E
σ(τ) exp τ dτ = ε(τ) exp τ | = ε(t) exp t − ε(0).
η η η 0 η
0

We rewrite this expression regarding strains:

( ) ∫t ( )
E 1 E
ε(t) = ε(0) exp − t + σ(τ) exp (τ − t) dτ. (6.12)
η η η
0

Thus, for the Kelvin model, the initial strain distribution fades exponentially over
time, and stresses are introduced as an integral convolution with the exponential
core.
Let’s consider a creep test with the creep load σk from the time t = 0 and
initial strains ε(0) = 0 (Fig. 6.7, Bertram, A. Solid Mechanics. Theory, Modeling,
and Problems/Albrecht Bertram, Rainer Glüge—Springer International Publishing
Switzerland, 2015. 331 p.). In accordance with (6.12), we obtain the following
110 6 Mathematical Models of Solid with Rheological Properties

Fig. 6.7 Creep tests within


the Kelvin model

expression:

∫t )
( ( )
1 E E
ε(t) = σk exp τ dτ exp − t
η η η
0
( )|t ( ) [ ( )]
1η E || E σk E
= σk exp τ exp − t = 1 − exp − t .
ηE η |0 η E η

From this expression, we get:


( )
. σk E
ε(t) = exp − t .
E η

Therefore, in this case there is no instantaneous elasticity. Creep speed slowly


decreases and creep strains converge to elastic strains: ε∞ = σk /E.
Relaxation time for the Kelvin-Foigt model tk = η/E is the time during which
strains increase to the value:
σk ( ) σk
ε(tk ) = 1 − e−1 ≈ 0.632 .
E E
The relation of the relaxation time to the value of initial strain rate ε̇(t) is: ε̇(0) =
ε∞ /tk .
6.2 Solid Mathematical Models 111

Relaxation test is not feasible for this model because limited (final) loads cannot
produce fracturing strains.
The constitutive equations for the Kelvin-Foigt medium can be written in the
following tensor representation:
( )
∂ei j 2 ∂θ
σi j = 2μei j + 2μtk + δi j λθ − μtk , (6.13)
∂t 3 ∂t

where δi j is Kronecker delta (δi j = 1, δi j = 0 under i /= j); tk is stress relaxation


time.

Remark If tk = 0, from (6.13) follows the Hooke’s law. Therefore, it focused more
on the shortcomings of the Maxwell and Kelvin-Foigt models.

• The Maxwell element has properties that contradict the real material properties,
namely: the invariability of strain rate during creep, the irreversibility of accumu-
lated strain component and the complete stress relaxation. The irreversibility of
strain component, which increases over time during creep and stress relaxation,
is the characteristic of elastic–plastic but not viscoelastic bodies.

The Kelvin-Foigt element does not have an elastic deformation component, so


that the initial value of the modulus of elasticity tends to infinity: E → ∞
Due to the marked negative properties, both considered elements are rarely used
as independent models.
We can obtain media with a variety of properties by combining Maxwell and
Kelvin-Foigt models (connecting mechanical analogues in series or in parallel). We’ll
consider some of the most commonly used models.
• Poynting model
The Poynting model (another name is the “standard” linear solid model) is the
“parallel connection” of an elastic body with a Maxwell body or the “consecutive”
connection of a Kelvin body with an elastic body (Figs. 6.3c and 6.8).
We perform a serial transformation in accordance with a similar procedure that
used for previous models:

Fig. 6.8 The Poynting model or the “standard” linear solid model
112 6 Mathematical Models of Solid with Rheological Properties

σ = σ K + σC = σ D + σC ; σC = CεC , σ K = K ε K , σ D = Dε̇ D

ε = εC = ε K + ε D , ε̇ = ε̇C = ε̇ K + ε̇ D .

Therefore: ε̇ = σ̇CC = σ̇KK + σDD = σ̇−σ̇


K
C
+ σ−σ
D
C
.
After performing further transformations, we get:

σ̇ C σ C
ε̇ = − ε̇ + − ε. (6.14)
K K D D
And based on (6.14):

D C+K
σ+ σ̇ = Cε + D ε̇. (6.15)
K K
As a result, we get the following exact expressions for the stress–strain state
component with initial conditions at t = 0 σ(0) = σ0 , ε(0) = 0 [Bertram, A. Solid
Mechanics. Theory, Modeling, and Problems/Albrecht Bertram, Rainer Glüge—
Springer International Publishing Switzerland, 2015. 331 p.]:
[ ∫ ( ) ] ( )
K2 t K K
σ(t) = (C + K )ε(t) + σ0 − ε(τ) exp τ dτ exp − t ,
D 0 D D
[ ( )
σ(t) σ0 K KC
ε(t) = + − + −
C+K C+K D(K + C) D(K + C)2

∫t ( ) ( )
KC ⎦ KC
σ(τ) exp τ dτ × exp − t . (6.16)
D(K + C) D(K + C)
0

The Poynting model demonstrates the properties of creep (Fig. 6.9a) and relaxation
(Fig. 6.9b) and instantaneous elasticity (Fig. 6.9a) [Bertram, A. Solid Mechanics.
Theory, Modeling, and Problems/Albrecht Bertram, Rainer Glüge—Springer Inter-
national Publishing Switzerland, 2015. 331 p.].
To describe the Poynting model, the following representations are often used:
( ) ( )
∂ Si j 1 ∂ei j 1 ∂σ 1 ∂e 1
a + Si j = 2μ + εi j ; a1 + σ = 3k + ε ,
∂t τ ∂t τ ∂t τ1 ∂t τ1

where εi j and Si j are the components of strain and stress deviators; μ and k are the
initial elastic shear modulus and volume expansion modulus; a, a1 are rheological
coefficients (a ≥ 0, a1 ≤ 1);τ, τ1 are shear and average stress relaxation times.
• Burgers model
Burgers model is a model formed by consecutively connecting a Maxwell body
and a Kelvin body or a parallel connection of two Maxwell bodies. Burgers model
6.2 Solid Mathematical Models 113

Fig. 6.9 Material behavior


corresponding to the
Poynting model: a is the
creep curve; b is the
relaxation curve

describes a medium with both elastic-lag and stress relaxation and creep properties
(Figs. 6.3d and 6.10). This model shows instantaneous elasticity due to the presence
of spring C.
Burgers model can be introduced as a parallel connection of two Maxwell models.
In accordance with this, we have:

σ = σC = σ D = σ K + σ R ; σC = CεC , σ K = K ε K , σ D = Dε̇ D , σ R = Rε̇ R ;


ε R = ε K , ε = εC + ε D + ε K ,R .

Therefore:

σ̇ = K ε̇ K + Rε̈ R = K (ε̇ − ε̇ .C − ε̇ D ) + R(ε̈ − ε̈C − ε̈ D )


( ) ( )
σ̇ σ σ̈ σ̇
= K ε̇− − + R ε̈ − − .
C D C D

Finally, we have:
( )
σ̈ 1 K 1 K K
+ + + σ̇ + σ = ε̈ + ε̇. (6.17)
C D CR R DR R

Fig. 6.10 Structural element


of the Burgers model
114 6 Mathematical Models of Solid with Rheological Properties

The numerical integration allows us to transform the second-order differential


Eq. (6.17) into a system of two first-order differential equations if we introduce the
stress in the spring K as an internal variable:

σ̇ σ σ σK K K
ε̇ = + + − and σ̇ K = K ε̇ − σ̇ − σ,
C D R R C D
or in stress increments:
( σ σ σK ) K
σ̇ = C ε̇ − − + , σ̇ K = (σ − σ K ).
D R R R
The creep curve for Burgers model under the creep load σ K (Fig. 6.11a) is the result
of the superposition of creep curves of the Kelvin and Maxwell models [Bertram,
A. Solid Mechanics. Theory, Modeling, and Problems/Albrecht Bertram, Rainer
Glüge—Springer International Publishing Switzerland, 2015. 331 p.]:
[ ( )]
σk K σk σk
ε(t) = 1 − exp − t + t+ . (6.18)
K R D C

At the beginning of the creep process, the damper R contributes to the creep
strains (initial creep). The contribution of the damper R decreases more and more
over time, and then the damper D begins to dominate with a steady creep speed σk /D.
This is called the second or steady-state creep phase. For many real materials, this
creep phase continues until the third creep phase, which cannot be described only
by Burgers model.

Fig. 6.11 Material behavior


corresponding to the Burgers
model: a creep curve; b
relaxation curve
6.2 Solid Mathematical Models 115

Fig. 6.12 Structural


diagram of the Burgers
model

In the relaxation test (Fig. 6.11b), the stresses in spring C are relaxed due to the
displacement of damper D and they tend to zero.
In general, the Burgers model can be described using this equation:

∂ei j ∂ 2 εi j ∂ Si j ∂ 2 Si j
a + b 2 = Si j + c + d 2 , a, b, c, d > 0.
∂t ∂t ∂t ∂t

Consider some of the Burgers model features in more detail [Bertram, A. Solid
Mechanics. Theory, Modeling, and Problems/Albrecht Bertram, Rainer Glüge—
Springer International Publishing Switzerland, 2015. 331 p.].
According to the Burgers structural scheme (Fig. 6.12), the relationship between
strain and stress can be described using a second-order differential equation in this
representation:
( )
η1 η2 η1 η2 η1 η2 η1 η2 η1 η2
ε̈ + η1 ε̇ = σ̈ + + + σ̇ + σ. (6.19)
E2 E1 E2 E2 E1 E1

Direct integration of Eq. (6.19) is difficult due to the instability of the numerical
method (second derivatives reach large values, which is caused by a strong oscillation
of a system). Therefore, from the point of view of numerical implementation, it seems
effective to implement the Burgers model at a uniaxial load, but the calculation is
stable.
The extension of the Burgers model at constant stress (creep) is expressed using
the creep function in the following form (Fig. 6.13) [Bertram, A. Solid Mechanics.
Theory, Modeling, and Problems/Albrecht Bertram, Rainer Glüge—Springer Inter-
national Publishing Switzerland, 2015. 331 p.]:
( ( ( )))
σ E1 E1 E2
ε= 1+ t+ 1 − exp − t . (6.20)
E1 η1 E2 η2

We show that the Burgers material creep at the beginning of the process corre-
sponds to the Kelvin material (damping), at the end of the process corresponds to
the Maxwell material (reaching the straight line of uniform extension).
We decompose the function (6.20) into a Taylor row when t → 0. As a result, we
obtain:
116 6 Mathematical Models of Solid with Rheological Properties

Fig. 6.13 Burgers model


extension at constant stress

( ( ( ))) ( )
σ E1 E1 E2 σ E1 E1
ε= 1+ t+ 1 − exp − t = 1+ t+ t + o(t) .
E1 η1 E2 η2 E1 η1 η2

For physical reasons, when t → 0, the “greater weight” has a viscous element η1 ,
that is, it can be considered that most often η1 >> η2 . For this reason:
( )
σ E1
ε≈ 1+ t .
E1 η2

This expression corresponds to the Kelvin creep function:


( ( ))
σ E2
ε= 1 − exp t .
E2 η2

Remark In this case, E 1 = E 2 is accepted, but this is not essential for demonstration.

Similarly, when t → ∞, we obtain:


( )
σ E1
ε≈ 1+ t .
E1 η1

That is, creep begins to correspond to Maxwell material. Therefore, in the Burgers
model, two components can be distinguished:
1. The first component corresponds to the Kelvin model (Fig. 6.14) and it is
described using the equation:

E2 σ
ε̇ + ε= .
η2 η2

It contributes significantly to the general expression only at high strain rates (ε̇).
2. The second component of the Burgers model corresponds to the Maxwell model
(Fig. 6.15) and it is described using the equation:
6.2 Solid Mathematical Models 117

Fig. 6.14 Kelvin model


extension graph

η1
η1 ε̇ = σ + σ̇. (6.21)
E

It makes a significant contribution at low strain rates (ε̇).


Under constant load σ (t) = σ1 , converting Eq. (6.21), we obtain:

E1
σ= ε. (6.22)
1 + (E 1 /η1 )t

The resulting expression is similar to the models of accumulated damage in the


elements (in our case in springs).
There are several approaches to construct algorithms for the numerical imple-
mentation of the “accumulated destruction” process.
The first approach is as follows: we record the time that the spring was under
loading, the value of which is greater than σ0 (we consider that with a lower load,
there is no creep). When calculating the bonding forces, the spring is considered
“degraded” and its stiffness is replaced as follows:

k
k→ .
1 + (k/η1 )t

Fig. 6.15 Maxwell model


extension graph
118 6 Mathematical Models of Solid with Rheological Properties

Table 6.1 Various


Hooke model P0 σ = Q0ε
viscoelastic differential
models Newton model P0 σ = Q 1 ε̇
Maxwell model P0 σ + P1 σ̇ = Q 1 ε̇
Kelvin model P0 σ = Q 0 ε + Q 1 ε̇
Poynting model P0 σ + P1 σ̇ = Q 0 ε + Q 1 ε̇
Burgers model P0 σ + P1 σ̇ + P2 σ̈ = Q 1 ε̇ + Q 2 ε̈
Model (p, q)-type P0 σ + · · · + Pp σ( p) = Q 0 ε + · · · + Q q ε(q)

Second approach (integral). We just integrate the expression (6.21). The result is
the following formula for stress increment:
⎛ ⎞
∫t
1
σ̇ = E 1 ε⎝1 − σdt ⎠.
εη1
0

Then the spring stiffness is changed as follows:


⎛ ⎞
∫t
1
k → k ⎝1 − σdt ⎠.
εη1
0

• Viscoelastic Differential Models

Table 6.1 summarizes the differential equations of the described rheological


models. These equations are generally presented in the form as ordinary differential
equations (ODE) of the order p with a time derivative of stresses and of the order q
with a time derivative of strains.
Constants P0 , . . . , Pp and Q 0 , . . . , Q q are determined by the properties of springs
and dampers. In all cases, one of the non-zero constants must be normalized using
the unit.
These differential equations fully describe the behavior of their respective models.
• Viscoelastic Models of Integral Type
The following models describing linear viscoelastic processes are based on the
Boltzmann superposition principle.
The Boltzmann superposition principle can be formulated as follows: if the effects
of deformation processes εi (t), i = 1, 2 are stresses σi (t), i = 1, 2, then the
superposition ε1 (t) + ε2 (t) leads to stresses σ1 (t) + σ2 (t).
We describe the procedure for obtaining a viscoelastic model of integral type.
6.2 Solid Mathematical Models 119

Let a constant load Δσi (t) act on the body during the period of time ti . This
loading process can be determined through the Heaviside1 function H (t) = 0, t <
0; H (t) = 1, t ≥ 0 in the following form:

σi (t) = Δσ(ti )H (t − ti ).

That is, the stress Δσ(ti ) remains constant from the time ti to the new time t j .
Let J (t − ti ) be the resulting deformation process from the action of a single load
H (t − ti ) that began at the time ti . According to the superposition principle, the load
Δσ(ti )H (t − ti ) initiates a deformation process of this type:

εi (t) = Δσ(ti )J (t − ti ).

The function J (t − ti ) is called the creep function.


An arbitrary loading process σ(t) can be approximated as the sum of the following
functions:

Σ
n
σ(t) = Δσ(ti )H (t − ti ).
i=1

Then the corresponding strains have the form:

Σ
n
ε(t) = Δσ(ti )J (t − ti ).
i=1

In the limit n → ∞, the strain function tends to the integral in the following form
(Stieltjes2 integral):

∫σ(t) ∫t
ε(t) = J (t − τ)dσ(τ) = J (t − τ)σ̇(τ)dτ. (6.23)
σ(0) 0

By changing the role of stresses and strains, we get this resulting equation [the
actions are the same as Eq. (6.23)]:

∫ε(t) ∫t
σ(t) = R(t − τ)dε(τ) = R(t − τ)ε̇(τ)dτ, (6.24)
ε(0) 0

1 Oliver Heaviside (1850–1925), an English mathematician and physicist who brought complex
numbers to circuit analysis, invented a new technique for solving differential equations (equivalent
to the Laplace transform), independently developed vector calculus.
2 Thomas Joannes Stieltjes (1856–1894), a Dutch mathematician. He was a pioneer in the field of

moment problems and contributed to the study of continued fractions.


120 6 Mathematical Models of Solid with Rheological Properties

where R(t) is the relaxation function.


The relaxation function and the creep function are related by the following relation
[follows from Eqs. (6.23) and (6.24)]:

∫t ∫t
t= R(t − τ)J (τ)dτ = J (t − τ)R(τ)dτ.
0 0

Remark Since all rheological models of the (p,q)-type satisfy the superposition
principle, all of them can be possibly represented with similar Stieltjes integrals.

All viscoelastic models of this type have the following characteristic features:
• They are linear.
• Their behavior is temporarily dependent that is viscous (rheonomic).
• They describe behavior with decaying memory. This means that if an event
occurred long enough ago, then its effect in the present is negligible.
• With very fast or very slow processes, these models behave elastically as long as
all dampers become rigid or pliable, respectively.
The rank of all these models is limited by their linearity. In general, linear
differential operators of various orders are used to describe linear viscous-elastic
media.
The Volterra3 principle is very important, which allows the results of solving
static problems in elasticity theory to be recalculated into the states of hereditary
viscoelasticity.

6.2.2 Creep Damage

A typical creep curve starts from the initial creep phase with high creep speeds.
Then the creep curve goes into the second creep phase with a minimum and
constant creep rate (steady creep phase). In the third stage, the creep rate increases
again (nonlinear creep phase), (Fig. 6.16) [Bertram, A. Solid Mechanics. Theory,
Modeling, and Problems/Albrecht Bertram, Rainer Glüge—Springer International
Publishing Switzerland, 2015. 331 p.]. The tertiary creep stage ends in destruction.
While the initial and second creep phases may be described using rheological
models within linear viscoelasticity, such models are not applicable to the tertiary
creep phase because they have non-linear behavior.
Non-linear behavior in the third phase of creep is a consequence of various mech-
anisms. Among them are such as a reduction in the transverse area, which leads to
an increase in real stresses; the appearance and growth of microcracks; changes of

3 Vito Volterra (1860–1940) was an Italian mathematician and physicist, known for his contributions
to mathematical biology and integral equations, being one of the founders of functional analysis.
6.2 Solid Mathematical Models 121

typical creep stages typical creep stages

Fig. 6.16 Typical creep curves with three creep phases

microstructure and others. The effect of these mechanisms is summed up into the
“Creep Damage” (or “Creep Rupture”) phase as a phenomenological theory.
An effective approach to describe the fracture of third creep stage was proposed
by Kachanov.4 The main idea of the Kachanov approach is that only a part of the
cross section of the sample S, namely a reduced cross section S l , withstands the force
when creeping. This cross section is called the effective cross-section. This leads to
the introduction of the concept of effective stresses:

σl = F/Sl . (6.25)

Effective stresses are greater than the true stresses σ = F/S.


If the effective stresses are introduced into the creep law, then the creep speed
increases in this case.
We enter the fracture parameter d, defined as follows:

S − Sl Sl
d= =1− or Sl = (1 − d)S. (6.26)
S S
Based on (6.26), it follows that the initial value of d is zero (d = 0) for the
undisturbed material and grows monotonically during creep. If the value of parameter
d approaches to one (d → 1), then effective stresses increase almost infinitely. In
fact, the creep process stops and the destruction occurs when the critical value dkr is
less than one: 0 < dkr < 1. The value dkr is considered as a material constant.
The failure parameter d is considered as an internal variable for which an evolu-
tionary equation is required. Therefore, for example, by analogy with the Burgers
model, you can use such equation:

4 Kachanov Lazar Markovich (1914–1993), Russian mechanic, the largest specialist in the field of
creep theory and strength theory.
122 6 Mathematical Models of Solid with Rheological Properties

ḋ = ασ̇l + βσl + γd,

where α, β, γ are constants.


Yu. N. Rabotnov proposed the following equation for the parameter d:
( )N
σl
ḋ = , (6.27)
μ

where μ, N are two positive constants.


This equation gives good results in many applications.
Equation (6.27) with (6.2) and (6.26) constitute the Kachanov-Rabotnov
destruction theory. According to the definition of effective stresses (6.25), the speed
of effective stresses is:
( )
d F Ḟ F
σ̇l = = − 2 Ṡl . (6.28)
dt Sl Sl Sl

In the case of a monotonous creep test, when the applied load F is constant, we
obtain Ḟ = 0.
The growth of the effective cross-sectional area according to (6.26) is:

Ṡl = (1 − d) Ṡ − ḋ S. (6.29)

It is often assumed that the creep strains (but not fracture strains) occur
isochorically (for example, this is a general assumption for metals).
In addition, if we neglect the change in volume at the elastic stage, then for strains
˙ 0 and current volume V = l S, we
ε, defined as ε = (l − l0 )/l0 , strain rate ε̇ = l/l
obtain that the volume change of the prismatic sample is:

V̇ = 0 = l˙S + l Ṡ = ε̇l0 S + l Ṡ.

From this, we obtain that:

ε̇l0 S ε̇
Ṡ = − =− S.
l 1+ε

Finally, the increase of effective stresses is:


[ ] [ ]
Ḟ F ε̇ Ḟ F ḋ ε̇
σ̇l = + 2 ḋ S + (1 − d) S = + + . (6.30)
Sl Sl 1+ε Sl Sl 1 − d 1+ε

Based on this, it can be concluded that the growth of effective stresses occurs due
to:
• Increments of applied load F;
6.2 Solid Mathematical Models 123

• Growth of failure parameter d;


• Transverse contraction.

6.2.3 General Comments

The “mechanical” setting of problems of creep theory is usually formulated as


follows: we study the state of solid for a certain time interval 0 ≤ τ ≤ t when
a given system of external loads and temperature acting on the body. The internal
interaction between isotropic particles is characterized by stress, strain, and strain
rate tensors. An equation or system of equations is used to connect the parameters of
stress–strain state; specify a basic experiment to define material constants; find out
the law of changing deformations over time.
f p
It is traditional to separate complete strains εi j into elastic εiej , plastic εi j and
viscous (creep strain) εicj components:

f p
εi j = εiej + εi j + εicj (i, j = 1, 2, 3), (6.31)

Remark In the case where these variables are used without indices, we assume that
they correspond to the longitudinal component of the same strain under uniaxial
loading.

Currently, two alternative ideas are mainly used in constructing the defining creep
equations.
The first idea is based on the concept of a mechanical equation of state. The
initial ratios for creep strain rates ε̇icj in this case are:

∂f ∂f
ε̇icj = λ + binj , (6.32)
∂σi j ∂qn

where f is creep potential given by the ratio:


[ ]
f = f I2 (Dσ ), I3 (Dσ ), ε̇ic , q1 , . . . , qr , T . (6.33)

As follows from expression (6.33), the creep potential includes the invariants
of the stress deviation I2 (Dσ ), I3 (Dσ ), the intensity of the creep strain rate ε̇ic and
some internal parameters qn , the kinetics of which are described using the following
evolutionary equation:

dqn = ainj dσi j + binj dεicj + cn dt + d n dT , (6.34)

where a n , bn , cn , d n are some functions from σi j , εi j , t, T , as well as from parameters


qn , (n = 1, r ), if the number of internal parameters is r.
124 6 Mathematical Models of Solid with Rheological Properties

The other designations in (6.32)–(6.34) are as follows: λ is proportionality


coefficient; t is time: T is temperature. Summation is based on duplicate indices.
Depending on the internal variable qn contained in the creep potential f, we obtain
one or another creep theory.
In particular, time t is∫ included in flow theory, and in hardening
∫ theory, the work
of plastic deformation σi j dεicj , the Odqvist parameter dεic or the value of the
accumulated creep strain εicj are included. In the theory of aging, instead of the
potential of creep rates, the potential of creep strains is used, instead of the rate of
creep strain intensity ε̇ic , the actual intensity of creep strains εic is used, and the time
t is set as an internal parameter.
To describe the third stage, internal parameters of scalar, vector, or tensor nature,
reflecting the process of damage accumulation are additionally included in the creep
potential f .
In the framework of flow theory, for example, for the simplest case of one-
dimensional creep, it is accepted:
( )n
dεc σ
=B , (6.35)
dt 1−ω

where ω is a scalar damage parameter for which it is generally possible to be written


as:
[ ]
dω σ(1 + εc ) m 1
=C ωβ , (6.36)
dt (1 − ω ) (1 − ω)q
r

where C, m, r, q, β are experimentally determined coefficients (C > 0; m ≥ 1;


q ≥ 0; β ≥ 0).
Another approach to construct defining creep equations is based on the use of
Volterra-Frechet5 integral representation:

e = J1 ∗ dσ + J2 ∗ dσdσ + · · · , (6.37)

In (6.37), it is accepted that e = εe + εc and Ji are the weight functions of time.


As simplified options (6.37), the following relationships are most often used:

∫t
ε = ψ(σ) + J0 (t1 , t2 , . . . , tn )σ(τ)dτ (6.38)
0

and

5 Maurice Rene Frechet (1878–1973), French mathematician.


6.2 Solid Mathematical Models 125

∫t
ϕ(ε) = σ + J0 (t1 , t2 , . . . , tn )σ(τ)dτ, (6.39)
0

where J0 is the Volterra integral operator with a special type of difference or non-
difference core; ψ is the stress function.
Equations (6.38) and (6.39) (unlike the first approach) take into account the
loading history, but the third creep stage is also described using the damage parameter.
The concretization of Eqs. (6.33), (6.38), and (6.39) is carried out on the basis
of certain concepts (for example, aging or hardening theories) implemented in
experiments (most often for uniaxial stretching).
Generalization in the case of a complex stress state and non-stationary loading
is carried out using the hypothesis of a “single creep curve”. Therefore, all the
limitations and disadvantages of the one-dimensional models in this approach are
completely preserved.
The most principle is the inadequacy of the description of the third creep stage.
For example, Eq. (6.35) together with (6.36) solve this problem approximately, since
the true law of the damage effect on creep speed is unknown.
When choosing a rheological equation for a specific material, the following
procedure must be followed:
1. Static and dynamic experiments are carried out with samples from the examining
material. For example, a diagram of uniaxial tension–compression is plotted at
static and dynamic loading (with different loading rates).
2. Constructed experimental curves are compared with theoretical creep curves.
By comparing the curves of the same name, a rheological body is selected, the
theoretical curves of which are most “close” to the corresponding experimental
curves. From the comparison of the same-name (theoretical and experimental)
curves, the values of rheological parameters for the defined rheological equation
are specified.
This path selected by the rheological equation is the most common. Therefore,
in the general case, the rheological equation describing the deformation process
can be represented by a functionality in the following form:

F(t, ε, ε̇, ε̈, . . . , s, ṡ, s̈) = 0, (6.40)

or in the form of functionals permitted relative to stress and strain components:


[( ) ] [ ( ) ]
t t
s(t) = F1 ε τ , t , ε(t) = F2 s τ , t . (6.41)
0 0

Functionals (6.41) are continuous in the region of functions s(t) and ε(t). If you
designate the required functions as z(t) and y(t), these functionals can be represented
in the following form:
126 6 Mathematical Models of Solid with Rheological Properties

∫t
az(t) = y(t) + b K 1 (t, τ1 )y(τ1 )dτ1
0

2 ∫t ∫t
b
+ K 2 (t, τ1 , τ2 )y(τ1 )y(τ2 )dτ1 dτ2 , (6.42)
2!
0 0
∫t ∫t
bn
··· + ... K n (t, τ1 , . . . τn )y(τ1 ) . . . y(τn )dτ1 . . . dτn
n!
0 0

where a, b are constants; K n (t, τ1 , . . . τn ) are known functions specified indepen-


dently from y(τ).
Remark This conclusion follows from the Weierstrass6 theorem that any continuous
function on the interval (c, d) can be approximated with a pre-set accuracy of the
usual power series.
Let’s limit in the right parts (6.42) to two terms and take into account that for “non-
aging” linear-hereditary media, the kernels of integral equations can be different.
Then they can be represented in the following form:
⎡ ⎤
∫t
1 ⎣
ε(t) = s(t) + J (t − τ)s(τ)dτ⎦, (6.43)

0
⎡ ⎤
∫t
s(t) = 2μ⎣ε(t) − R(t − τ)ε(τ)dτ⎦. (6.44)
0

The representations for ε(t) and s(t) in the form (6.43) and (6.44) are very conve-
nient, since they make it possible to move away from mechanical structural models
and introduce new abstract models of rheological bodies—mathematical models of
rheological bodies.
Note that the expressions (6.43) and (6.44) are general, si j and εi j are components
of stress and strain deviators, i.e., si j = σi j − σ · δi j , εi j = ei j − δi j e.
By setting functions J (x, t) and R(x, t), and associating a new rheological body
with them, an infinite set of new mathematical models of rheological bodies can be
obtained. All these mathematical models will be linear models. They characterize
the stress–strain state of hereditary elastic bodies.
Representation (6.44) can be considered as the solution of (6.43). Therefore,
R(t, τ) in (6.44) is resolvent, J (t, τ) in (6.43) and vice versa.
The expression (6.43) can be divided into two parts:

6Karl Theodor Wilhelm Weierstrass (1815–1897), German mathematician often cited as the "father
of modern analysis".
6.2 Solid Mathematical Models 127

• The first term is instantaneous deformation (elastic)

εel (t) = s(t)/2μ;

• The second term is hereditary deformation (also elastic):

∫t
1
ε (t) =
sp
J (t − τ)s(τ)dτ,

0

where the hereditary function J (t − τ) “remembers” the effect of loading at the time
τ on the body deformation at the moment of observation t.
According to the physical meaning J (t − τ) must be decreasing function and in
case of μ = constant, it must be function of not t and τ, but (t − τ), since it depends
only on the time passed from the moment τ of loading beginning to the moment t
when we observe deformations from loading.
If on the right side of the expression (6.42), we leave respectively 3, 4, 5 … first
terms, then we get the non-linear integral equations respectively 3, 4, 5th … orders
that can be mathematical models of the corresponding rheological bodies.

Remark It should be emphasized that such a way to build new mathematical models
of rheological bodies is not the only approach.

For rheological equations in which the coefficients are constant, the kernels
K (x, t) and ζ(x, t) in the corresponding integral equations can be represented as
exponential functions, i.e., as follows:

Σ
n
( ) ( )
exp −k j (t − τ) , k j > 0 , ( j = 1, 2, . . . , n). (6.45)
j

With the complication of integral Eq. (6.42) and the inclusion of a larger number
of terms, there are great mathematical difficulties in solving them. Therefore, in this
case, instead of (6.42), non-linear integral equations of the following form are more
often considered:

∫t
ϕ(ε) = s(t) + J (t − τ)s(τ)dτ, (6.46)
0

or such:
⎡ ⎤
∫t
1 ⎣
ε(t) = s(t) + J (t − τ) f [s(τ)]dτ⎦, (6.47)

0
128 6 Mathematical Models of Solid with Rheological Properties

where ϕ(ε) and f [s(t)] are defined functions characterizing non-linear deformations.

6.3 Rheological Equations of Linear Viscoelasticity

6.3.1 General Concepts

First, recall some common basic concepts introduced earlier.


Physical relationships between stress and strain deviators (si j = σi j − σ · δi j ,
εi j = ei j − δi j e) in the case of the linear viscoelasticity have the following form:

∫t
2μεi j (t) = si j (t) + J (t − τ)si j (τ)dτ, K θ(t) = σ(t), (6.48)
0

where θ = ekk is the relative volume change; σ = σkk /3 is the average (hydrostatic)
stress; μ is the instantaneous elastic shear modulus; K is the instantaneous volume
deformation modulus and function J(t) is the creep kernel.
Physical interpretation of Eq. (6.48): the field of deformations εi j at a current
moment of time is determined not only by the instantaneous stress si j (connect with
strains by the generalized Hooke’s law), but also by previous stress values by means
of some hereditary functions. The volumetric strain θ is taken to be elastic since the
volumetric creep is small compared to the shear one.
Equation (6.48) is invariant with respect to the beginning of the time reference,
since the creep kernel J(t) has a difference (t – τ ) with its arguments. The kernel
J(t) is a monotonically positive decreasing function of its arguments. At infinity, it
asymptotically tends to zero. It is identical to zero, if the argument is negative.
If Eq. (6.48) can be resolved with respect to si j and σ, then:
⎛ ⎞
∫t
si j (t) = 2μ⎝εi j (t) − R(t − τ)εi j (τ)dτ⎠, K θ(t) = σ(t). (6.49)
0

Function R(t) is the relaxation kernel. It is a kernel J(t) resolvent. Of course, the
function J(t) in turn is the resolute kernel R(t) resolvent:

∫t
R(t) − J (t) = J (t − τ)R(τ)dτ. (6.50)
0

Deviator ratios (6.48), (6.49) can be written in another form:


6.3 Rheological Equations of Linear Viscoelasticity 129

∫t ∫t
εi j (t) = K (t − τ)dsi j (τ)dτ, si j (t) = F(t − τ)dεi j (τ)dτ, (6.51)
0 0

where K(t) is the creep function, F(t) is the relaxation function.


Between these functions and the previously introduced kernels J(t), R(t), a
connection can be established by integrating the ratios (6.51) in parts:

∫t
|t dK
εi j (t) = K (t − τ) si j (τ)|0 + si j (τ)dτ.

0

We accept sij (0) = 0 and K (0) = 1/(2μ), J (t) = 2μdK (t)/dt. As a result, we
come to Eq. (6.48). Similarly, the identity of Eq. (6.49) and the second equation of
(6.51) can be established.

6.3.2 Examples of Creep and Relaxation Kernels

The functions J(t) and R(t) are defined most often as the results of experiments on
pure shear stress test.
Based on the experiments, creep curves are built ε12 (t)/σ12 0 ~ t. In this case,
all other components of stress and strain tensors are zeros. Equation (6.48) will be
reduced to one:

∫t
2με12 (t) = s12 (t) + J (t − τ)s12 (τ)dτ.
0

From this expression, in the case of creep at constant stress σ12 (t) = σ12
0
= const,
we obtain:

∫t
ε12 (t)
2μ 0 = 1 + J (ξ)dξ.
s12
0

The obtained expression allows you to determine the function J(t):

dK (t) ε12 (t)


J (t) = 2μ , K (t) = 0 .
dt s12

The K(t) is the pliability (creep function) in pure shear. Similarly, relaxation
kernel R(t) can be determined by the pure shearing experiments. Various analytical
130 6 Mathematical Models of Solid with Rheological Properties

expressions for creep and relaxation kernels are used to approximate experimental
data.
The simplest kernel of the integral equation of the most commonly used linear
hereditary creep theory is the whole exponent. However, we can describe real creep
curves with only one exponential function and only in a short time interval. Therefore,
kernels as the sum of exponent are more often used.
The main disadvantage of exponential kernels is that a large number of material
parameters are required to describe the mechanical behavior of real materials. In
addition, for such kernels, the strain rate dε/dt at the initial time t = 0 is a finite
value, but numerous experiments indicate that the strain rate at the initial time is
large, and its concrete determination is very difficult. Therefore, the heredity kernels
are usually given using various weakly singular functions that set an infinitely large
creep rate at t = 0.
1. Exponential creep kernel:

J (t) = Ae− pt (A > 0, p > 0).

Unfortunately, there are only two material constants A and p in this representation,
making it difficult to describe the experimental data over a wide time range, especially
in the initial loading period.
The resolvent kernel (relaxation kernel) has the form:

R(t) = Be−qt .

Substituting the exponential creep kernel and its resolvent into the ratio (6.50)
and requiring that it be an identity by t, we obtain:

B = A, q = A + p.

A more accurate representation gives the kernel and its resolvent as a sum of
exponential functions:

Σ
m Σ
m
J (t) = Ai e− pi t , R(t) = Bi e−qi t .
i=1 i−1

Substituting these expressions into the integral relationship between creep and
relaxation kernels (6.50) and requiring them to be identical, we obtain 2m equations
connecting the constants of resolvents Bi and qi and the kernel constants Ai and pi :

Σ
m
Ai Σm
Bi ( )
= 1, = 1, j = 1, m
i−1
q j − pi q
i−1 i
− p j
6.3 Rheological Equations of Linear Viscoelasticity 131

These equations are useful in determining one of the kernels if the others are
known. For example, if we determined the constants Ai and pi after processing exper-
imental data. Then, from the first system of equations, the roots qi are determined.
They must be real, unequal among themselves and different from pi . After that,
constants Bi are determined by the second system of linear equations.
2. Power kernel Duffing7 :

C
J (t) = (0 < β < 1).
t (1−β)
The limitation on the constant β is due to the fact that the strain rate and the strain
itself at the time of the loading application become infinitely large when β = 0.
The Duffing kernel “works worse” with sufficiently long loading periods.
The Duffing kernel resolvent is the following function:

Σ C k Γ∗k (β)t kβ−1


R(t) = (−1)k+1 .
Γ∗ (kβ)

where Γ∗ (β) is the gamma-function of the argument β.


With limitations on the value of β (0 < β < 1) this series converges everywhere.
This type of kernel is called kernel with a weak singularity or kernel with a weak
feature. This feature is that at t → 0, J(t) → ∞. This corresponds to an infinitely
high strain rate at the moment of the loading application. A weak desire for infinity
is provided by the restriction 0 < β < 1.
The Duffing kernel cannot satisfactorily describe the experimental data with a
wide time interval due to the presence of a limited number of arbitrary constants.
3. Rzhanitsyn8 core:

Ae− p(t−τ)
R(t − τ) = , ( p > 0, 0 < β < 1). (6.52)
(t − τ)(1−β)

This is a typical relaxation core in which exponential and weakly singular


properties are combined.
Kernel (6.52) resolvent is the following function:

e− p(t−τ) Σ Ak Γ∗k (β)(t − τ)kβ
J (t − τ) = (−1)k+1 ,
t − τ k=1 Γ∗ (kβ)

where Γ∗ (β) is the gamma-function.

7 Georg Wilhelm Christian Caspar Duffing (1861–1944), German physics-engineer.


8 Aleksey R. Rzhanitsyn (1911–1987), Soviet famous scientist, mechanic.
132 6 Mathematical Models of Solid with Rheological Properties

It differs from the Duffing core resolvent only by the multiplier e−pt . The
Rzhanitsyn core is often used in applied research.
4. Koltunov9 kernel.
Koltunov kernel has a more general representation than the Rzhanitsyn kernel:
α
Ae− p(t−τ)
R(t − τ) = , ( p > 0, 0 < α < 1, 0 < β < 1) (6.53)
(t − τ)(1−β)

The resolvent of this kernel is the function:


[ ( )]k
α+β−1
e− p(t−τ)α ∞
Σ AΓ∗ α (t − τ)k(α+β−1)
J (t − τ) = (−1)k+1 ( ) .
(t − τ)α k=1 Γ∗ k α+β−1
α

5. Abel10 kernel:

(t − τ)α
Jα (t − τ) = (−1 < α < 0), (6.54)
Γ∗ (1 + α)

where Γ∗ (1 + α) = αΓ(α) is the gamma–function.


Abel kernel allows the reversibility of integral ratio (6.48), wherein the constant
α is determined by experimental creep or relaxation curves. Abel kernel corresponds
to experimental data only in the initial stage of deformation, but then the unlimited
growth of deformation is not confirmed by the experiment. The applicability of Abel
kernel can be substantially extended by moving to the non-linear hereditary elasticity
equations.
6. Rabotnov fractional-exponential kernel.
The Rabotnov function is a generalization of the Abel kernel (for the first time
for β = –1):

Σ βn (t − τ)n(1+α)
.α (β, t − τ) = (t − τ)α .
n=0
Γ∗ [(n + 1)(1 + α)]

Series .α (β, t − τ) converges for all values t and β.


Obviously: .α (β, t − τ) = Jα (t − τ).

9 Michail A. Koltunov, Soviet scientist-mechanic.


10 Niels Henrik Abel (1802–1829), Norwegian mathematician who made pioneering contributions
in a variety of fields. His most famous single result is the first complete proof demonstrating the
impossibility of solving the general quintic equation in radicals.
6.3 Rheological Equations of Linear Viscoelasticity 133

The Yu. N. Rabotnov function .α (β, t − τ) is called a fractional-exponential


kernel.
This core has become widespread in the literature, due to the great possibility of
describing the rheological properties of materials and the presence of a sufficiently
developed special algebra of the corresponding operators. At the same time, it is very
difficult to calculate the Rabotnov kernel, due to the presence of a large number of
defined constants.
7. Power core of general form.
In this case, the creep function is:

J (t − τ) = (t − τ)α−1 e−β(t−τ) . (6.55)

The physical meaning (6.55) follows from the following formula for the creep
function:

J (t) = (E/σ)ė(t).

Therefore, J (t − τ), with accuracy to a constant factor, is a strain rate function


of a rod that is in a uniaxial tension (or compression) state by stresses of a constant
value: σ = constant.
8. Wulfson11 -Koltunov kernel is a generalization of the Rzhanitsyn and Rabotnov
kernels:

Σ∞
An (t − τ)n(1+α)+α
K (A, α, β, t − τ) = a exp[−β(t − τ)] . (6.56)
Γ [(n + 1)(1 + α)]
n=0 ∗

9. The method of constructing linear defining equations, based on the use of the
Gavrillac-Negami ratio for the complex modulus of elasticity is promising (for
example, for composite, anisotropic materials):

[ ]−γ
E(i ω) = E 0 − (E 0 − E ∞ ) 1 + (i ωτ)(α+1) , 0 < γ ≤ 1,

where E 0 is the instantaneous modulus of elasticity and E ∞ is the long-term modulus


(about ω → 0).
Expressions for relaxation functions include the corresponding Gavrillac-
Negami kernel:
( )−γ
Γ α,β,γ ( p) = p (α+1) + β .

11 Iosif I. Wulfson, Soviet scientist-mechanic.


134 6 Mathematical Models of Solid with Rheological Properties

In general, many types of different heredity kernels are proposed. The advantages
of one or another kernel depend on the physical properties of the material and the
type of task to be solved. The choice of linear kernel is explained primarily by the
desire to simplify the integral equations of viscoelasticity, which in this case allow
reversibility, that is, an analytical solution to the problem of resolvent can be obtained.
However, it should be noted that the problem of obtaining new kernels continues
to be urgent.

6.3.3 Volterra Principle

Let’s describe more detail about the Volterra method related to the linear theory of
solving viscoelastic problems of the type (6.49).
Expressions (6.49) are rewritten in symbolic form by entering the following
integral operator G* :

si j = 2G ∗ .i j , σ = K θ, (6.57)

where

∫t
( )
G = G 1 − R∗ , R∗ f =

R(t − τ) f (τ)dτ, G is shear modulus.
0

By comparing (6.57) with the relations of Hooke’s law for the linear theory of
elasticity, it can be concluded that the physical relations of linear viscoelasticity have
exactly the same form as the equations of generalized Hooke’s law, and only the shear
modulus G is replaced by the operator G* .
In the operator form, we also represent Eq. (6.48) in the following form:

1
.i j = si j , σ = K θ,
2G ∗

where the operator 1/G* is:

∫t
1 1( )
= 1 + J∗ , J∗ f = J (t − τ) f (τ)dτ.
G∗ G
0

Substituting where the expression from (6.57) G ∗ = G(1 − R ∗ ), we get:

1
= 1 + J ∗.
1 − R∗
6.3 Rheological Equations of Linear Viscoelasticity 135

Thus, we determined the operation of dividing by some operators.


From where, we come to the following formula of the Voltérra principle: the
solution of linear problem of viscoelasticity can be obtained as solution of the corre-
sponding problems of linear theory of elasticity, in which elastic constants (proper-
ties) are replaced by some operators. In this case, the shear modulus G should be
replaced by operator G* .
Solutions to the problems of elastic theory, in addition to the shear modulus,
contain other elastic constants, for example, the modulus of elasticity E and the
Poisson’s ratio ν. In accordance with the described approach, in the elastic solution,
replace the modulus of elasticity E and Poisson’s ratio v with constants G and K, and
then replace the constant G with an operator G* .
Using the Voltérra principle, we get a solution including algebraic or transcendent
functions of time operators. In addition, this solution must also be “decrypted”. In
the general case, for example, the Laplace-Carson12 integral transform, the Ilyushin
approximation method, the Rabotnov operators, and others are used.
In particular, the Laplace-Carson image of the function f (t) is called the following
function f * (p) of the real parameter p:

∫∞

f ( p) = p f (t)e− pt dt.
0

If f (t) = f 0 , then f * (p) = f 0 . If the function f (t) is a convolution of the two


functions F and ψ, that is:

∫t
f (t) = F(t − τ)dψ(τ), (6.58)
0

then its image equals to the product of images of sub-integral functions:

f ∗ ( p) = F ∗ ( p)ψ∗ ( p). (6.59)

In turn, if the image of the function f (t) is determined by the relation (6.59), then
the function itself is represented by the formula (6.58).
The complete system of equations includes the equations of linear viscoelas-
ticity (6.51) with elastic volumetric strain, equilibrium equations, Cauchy ratios,
and boundary conditions:

∫t
εi j (t) = J (t − τ)dsi j (τ)dτ, σ = K θ; σi j, j + ρFi = 0;
0
( )
ei j (t) = u i, j + u j,i /2; σi j l j = Ri on Sσ , u i = u 0i on Su . (6.60)

12 J. Carson Mark (1913–1997), Canadian-American mathematician.


136 6 Mathematical Models of Solid with Rheological Properties

We apply the Laplace-Carson transform to the ratios (6.60) and considering (6.59),
we obtain:

.i∗j = J ∗ si∗j , σ∗ = K θ∗ , σi∗j, j + ρFi∗ = 0,


2ei∗j = u i,∗ j + u ∗j,i , σi∗j l j = Ri∗ on Sσ , u i = u ∗0i on Su . (6.61)

It follows that the linear viscoelastic problem in images is identical to the corre-
sponding problem for an elastic body. Therefore, the solutions to these problems
coincide with the only difference that the constant (1/2)G is replaced by the image
G* . The modulus of volume compression in the solution is saved. If we find the
solution of the linear problem of elasticity theory for images σi∗j , ei∗j , then we will
find the originals from these images and determine the solutions σi j (x, t), ei j (x, t).
Therefore, when solving problems for viscoelastic bodies, it is enough to first deter-
mine the solution of the same problem in the classical theory of elasticity, and then
in the found solution replace the elastic constants E and ν, operators E and ν.

Remark It should be noted that the Volterra principle is not the only method for
solving viscoelastic problems.

There is no exact classification of problems that can be resolved using the Volterra
principle. Therefore, it has been more or less established that for problems with
time-interchangeable boundaries, the Volterra principle is not suitable.
According to the Volterra principle, it follows that in all problems whose final
results do not include constant materials, the results of the classical theory of elasticity
can be right for any rheological body with the general law of linear deformation:
⎡ ⎤
∫t
si j (t) = 2μ⎣εi j (t) − ϕ(t − τ)εi j (τ)dτ⎦. (6.62)
0

Such problems include, for example:


1. The first main problem of the theory of elasticity in stresses for single-connected
regions (the problem of the stress state of the elastic body at external forces set
on its surface). In the formulas for stress components (rather than strains), the
constants of elastic material do not enter. Therefore, this solution can be right for
the rheological body with the general law of linear deformation (6.62).
2. Stress problems when examining the stress concentration around various holes
whose contour is free from external forces, or when the principal vector of
external forces applied to the hole contour is zero.
Solutions of viscous-elastic non-linear problems on stress concentration around
curvilinear holes show that the formulas for stress components near arbitrary holes
include elastic constants of non-linear materials. Thus, for viscoelastic materials,
not only the displacements, but also the stresses near the hole change over time.
In order to adequately describe the changes in stresses near arbitrary holes in real
6.4 Equations of Mechanics of Linear-Hereditary Media 137

materials without leaving the area of linear problems, it is enough to refuse the ideal
isotropic model and take as the main anisotropic model. In this case, it is possible to
use appropriate solutions to the problems of the theory of elasticity of an anisotropic
medium. Next, the Volterra principle is used, that is, instead of elastic constants,
operators of the form (6.57) are introduced and we obtain the law of changing stresses
near holes over time.

6.4 Equations of Mechanics of Linear-Hereditary Media

The method of obtaining a resolving system of equations of mechanics of small


viscoelastic strains for media subjects to the general law of linear deformation does
not differ from the method of obtaining these equations used in the classical theory
of elasticity. The only difference among these systems of equations is that the law
describing the relations between stresses and strains in the theory of elasticity must
be replaced by the corresponding rheological equations.
The complete system of equations of the theory of linear viscoelasticity in the
case of small strains has the form:
(a) Equilibrium equations
Σ ∂σi j
+ X i = 0, i = 1, 3 (6.63)
j
∂x j

(b) Boundary conditions


Σ ( )
σi j cos n, x j = X ni , (6.64)

(c) Saint-Venan strain compatibility equation

∂ 2 e11 ∂ 2 e22 ∂ 2 e12


+ =2 ,··· (6.65)
∂ x22
∂ x12 ∂ x1 ∂ x2

(d) Cauchy relation


( )
1 ∂u i ∂u j
ei j = + , (6.66)
2 ∂x j ∂ xi

(e) Rheological equations

Ee11 = σ11 − ν(σ22 + σ33 )


Ee22 = σ22 − ν(σ11 + σ33 )
138 6 Mathematical Models of Solid with Rheological Properties

Ee33 = σ33 − ν(σ11 + σ22 )


E E E
σ23 = e23 , σ13 = e13 , σ12 = e12 , (6.67)
1+ν 1+ν 1+ν

where
− ( ∗) ( ∗
)
E = E 1− E , ν =ν 1+ν , (6.68)

∫t ∫t
∗ ∗
E f (t) = ε(t − τ ) f (τ )dτ, ν f (t) = N (t − τ ) f (τ )dτ, (6.69)
0 0

where E, ν are respectively, modulus of elasticity and Poisson’s ratio at elastic


deformation of the material. ε(t − τ) and N (t − τ ) are heredity kernels.
From expressions (6.67), it follows that:
( )
− μ 3λ + 2μ λ
E= ; ν= ( ),
λ+μ 2 λ+μ

or

Eν E 2 E
λ= ; μ=G= ;χ = λ + μ = ; (6.70)
(1 + ν)(1 − 2ν) 2(1 + ν) 3 3(1 − 2ν)

where χ is the modulus of material volume deformation.


The introduction of two independent kernels ε(t − τ ) and N (t − τ ) [in accor-
dance with (6.69)] gives more opportunities to select analytical dependencies to
adequately describe the real material behavior.
Many materials (especially metals) at volumetric deformation follow the linear
elastic law, i.e., the Hooke’s law. For such materials, the characteristic χ = 3(1−2ν)
E

must be a constant, i.e., time independent. Therefore, for such materials χ = χ =


constant or

E E
= . (6.71)
3(1 − 2ν) 3(1 − 2ν)

Then
[ ] [ ]
( ( ∗
)) 1 1 − 2ν 1 − 2ν ∗
E = E 1− E ν= 1− E =ν 1+ E . (6.72)
2 E 2ν

In this case, we obtain [on the basis of (6.52), (6.53) and (6.60)] that only one
operator will be included in the formulas for stresses and strains of isotropic materials
whose volumetric deformation is linear elastic:
6.4 Equations of Mechanics of Linear-Hereditary Media 139
( ∗)
E = E 1− E , (6.73)

∗ ∫t
where E ·1 = ε(t − τ ) · 1 · dτ .
0
That is, in this case, the stress and strain components will have one function of
influence ε(t − τ ) .
The system (6.63)–(6.67) is a full system of equations of mechanics of linear-
hereditary media.
The hereditary theory of viscoelasticity corresponds to the bodies in which
mechanical structural models, the viscous element is connected in a pair parallel to
the elasticity element. Therefore, for example, Maxwell’s elastic-viscous medium,
Burger’s body and the like cannot be the basis of the viscoelastic hereditary theory,
since they describe inelastic deformations.
Therefore, functions R(t − τ) and J (t − τ) in mathematical models of the corre-
sponding rheological bodies in the hereditary theory of viscoelasticity should reflect
linear reversible processes.

Control Questions

1. With creep strains over time:


1) increase;
2) decrease;
3) do not change.
2. Stresses during relaxation over time:
1) increase;
2) decrease;
3) do not change.
3. Physical relationships of linear viscoelasticity:
[ ]
∫t
1) si j (t) = 2μ εi j (t) − R(t − τ)εi j (τ)dτ ;
0
2) 2Gϕ(εu ) .i j (t) = si j (t);
3) 2G .i j (t) = si j (t).
4. Volumetric deformation of viscoelastic bodies:
1) elastic;
2) plastic;
3) linearly viscoelastic.
5. Rheological properties of the material characterize:
1) modulus of volume strains;
2) plasticity function;
3) creep and relaxation kernels.
140 6 Mathematical Models of Solid with Rheological Properties

6. Creep and relaxation cores are connected by dependency:


1) linear;
2) integral;
3) differential.
7. In the exponential creep kernel the material constants:
1) one;
2) two;
3) three.
8. In the Duffing creep kernel, the material constants:
1) one;
2) two;
3) three.
9. Constant of material in Rzhanitsyn creep kernel:
1) one;
2) two;
3) three.
10. With increasing strain rate, the yield strength:
1) does not change;
2) increases;
3) decreases.
11. Volterra principle: the solution of the viscoelastic problem can be obtained from
the solution of the corresponding problem:
1) elasticity;
2) plasticity;
3) damage.
12. Maxwell model, Kelvin-Foigt model?
1) ε̇ = Eσ̇ + λσ ;
2) σ = Eε + λε̇.

Control Exercises
Remark
Exercises are taken from the book [Mase, G. T. Continuum mechanics for engi-
neers/G. T. Mase, G. E. Mase—CRC Press LLC, 1999. 380 p.]. We recommend
solving other exercises given in the book.

1. A four-parametric model consists of a Kelvin structural model and a Maxwell


model (Fig. E.1).
6.4 Equations of Mechanics of Linear-Hereditary Media 141

Fig. E.1 Four-parametric


structural model

Knowing that γMODEL = γKELVIN + γMAXWELL , where γ is strain, and taking


into account the operator equation (∂t ≡ ∂/∂t ) for Kelvin structural model (a)
σ12 = (G + η∂t )γ12 and for Maxwell model (b) (∂t + 1/τ)σ12 = (G∂t )γ12

Construct a relation describing the behavior of this model.


Answer: G 2 η1 γ̈ + G 1 G 2 γ̇ = η1 σ̈ + (G 1 + G 2 + η1 /τ2 )σ̇ + (G 1 /τ2 )σ .
2. Construct defining equations for the three-parametric models shown in Fig. E.2.
Answer:
a) γ̈ + γ̇/τ1 = [(η1 + η2 )/η1 η2 ]σ̇ + (1/τ1 η2 )σ;
b) σ̇ + σ/τ2 = (G 2 + G 1 )γ̇ + (G 1 /τ2 )γ;
c) σ̇ + σ/τ2 = η1 γ̈ + (G 2 + η1 /τ2 )γ̇.
3. The model consists of a Kelvin unit connected in parallel to the Maxwell unit,
(Fig. E.3).
Construct a defining relation for such a model.
Answer: σ̇ + σ/τ2 = η1 γ̈ + (G 1 + G 2 + η1 /τ2 )γ̇ + (G 1 /τ2 )γ.
4. For the four-parametric model shown in Fig. E.4, to define:
a) Defining relation;
b) Relaxation function G(t).
Remark
In this case G(t) is the sum of the functions G(t) of all parallel connected
structural units.
Answer:
a) σ̇ + σ/τ2 = η3 γ̈ + (G 1 + G 2 + η3 /τ2 )γ̇ + (G 1 /τ2 )γ;
b) G(t) = G 1 + G 2 exp(−t/τ2 ) + η3 δ(t).
142 6 Mathematical Models of Solid with Rheological Properties

Fig. E.2 Different types of


three-parametric structural
models

Fig. E.3 Example of a


Kelvin unit connected in
parallel to the Maxwell unit

Fig. E.4 Special type of


four-parametric structural
model

5. There is the history of the model stress over time in Fig. E.5.

Determine strains γ(t) for this load in the time interval:


a) 0 ≤ t/τ ≤ 2; b) 0 ≤ t/τ ≤ 4.
To get an answer in (b), use the superposition principle.
Answer:
6.4 Equations of Mechanics of Linear-Hereditary Media 143

Fig. E.5 One of a possible type of three-parametric structural models

a) γ(t) = σ0 J (2 − exp(−t/τ))U (t);


b) γ(t) = σ0 J (2 − exp(−t/τ))U (t) − σ0 J (2 − exp(−(t − 2τ)/τ))U (t − 2τ).
6. For the model shown in Fig. E.6, define:
a) Defining equation;
b) Relaxation function G(t);
c) Stresses σ(t) for the interval 0 ≤ t ≤ t1 , when strains are set by the
corresponding graph
Answer:

a) σ̇ + σ/τ = ηγ̈ + 3G γ̇ + (G/τ)γ;


b) G(t) = ηδ(t) + G(1 + exp(−t/τ));
c) σ(t) = λ(2η − η exp(−t/τ) + Gt)U (t).
7. For the model from the previous exercise, define the stresses σ(t) for the case
where the strains γ(t) are set by the graph shown in Fig. E.7.
Answer:
σ(t) = γ0 (ηδ(t) + G(1 + exp(−t/τ)))U (t) +
λ(η(2 − exp(−t/τ)) + Gt)U (t).

Fig. E.6 Four-parametric structural model and its graph of strain mode
144 6 Mathematical Models of Solid with Rheological Properties

Fig. E.7 The graph of


strains mode

8. The three-parameter model shown in Fig. E.8 undergoes the strain mode shown
in the graph (Fig. E.8).
Using the superposition principle, obtain an expression for the stress σ(t)
for the interval t ≥ t1 based on the expression for σ(t) for the interval t ≤ t1 .
Accept that γ0 /t1 = 2.
Answer:
for t ≤ t1 : σ(t) = λ(η1 (1 − exp(−t/τ1 )) + G 2 t)U (t);
for t > t1 :
σ(t) = λ(η1 (1 − exp(−t1 /τ1 )) + G 2 t1 )U (t1 )
+ λ(η1 (1 − exp(−(t − t1 )/τ1 )) + G 2 (t − t1 ))U (t − t1 ).
9. For the model shown in Fig. E.9, define the stresses σ(t) at: (a) t = t1 ; (b)
t = 2t1 and (c) t = 3t1 , for the case where the applied strains are given by
the corresponding diagram. For options (b) and (c) you use the superposition
principle.
Answer:

a) σ(t1 ) = (γ0 η/t1 )(2 − exp(−t1 /τ));


b) σ(2t1 ) = (γ0 η/t1 )(−2 + 2 exp(−t1 /τ) − exp(−2t1 /τ));
c) σ(3t1 ) = (γ0 η/t1 )(− exp(−t1 /τ) + 2 exp(−2t1 /τ) − exp(−3t1 /τ)).

Fig. E.8 Three-parametric structural model and its graph of strain mode for exercise E.8

Fig. E.9 Three-parametric structural model and its graph of strain mode for exercise E.9
6.4 Equations of Mechanics of Linear-Hereditary Media 145

Fig. E.10 Three-parametric structural model and diagram of strains develop over time

10. Define the stress function σ(t) for the model in Fig. E.10 for the case
where strains develop over time in accordance with the diagram γ(t) =
(γ0 /2G)(2 − exp(−t/2τ))U (t).
Answer: σ(t) = γ0 U (t).
11. The stress relaxation function is given by G(t) = a(b/t)m , where a, b and m
are constant, t is time. ( )m
Show that the creep function in this case is J (t) = a m1 π sin mπ bt under
m <1.
Remark
Use the identity for Laplace transformant: G(t)J (s) = 1/s 2 .
12. The three-parameter solid model is shown in Fig. E.11. Get defining relations
for such a model. Using these equations, define: (a) the relaxation function and
(b) the creep function for this model.
Answer:
( )
a) G(t) = G 2 + G 1 exp −t/τ(1 ; ) ( ( ))
b) J (t) = (1/(G 1 + G 2 )) exp −t/τ∗1 + (1/(G 2 )) 1 − exp −t/τ∗1 , τ∗1 =
(G 1 + G 2 )τ1 /G 2
13. The material model is shown in Fig. E.12.

a) Define a relaxation function for a given material G(t).


b) If the slope of the time strain change function is λ, determine the stresses
using the corresponding hereditary integral with G(t).

Answer:

a) G(t) = G + 2G exp(−2t/τ) + G exp(−t/2τ);


b) σ(t) = Gλ(t + 3τ − τ exp(−2t/τ) − 2τ exp(−t/2τ))U (t).

Fig. E.11 Three-parametric


structural model
146 6 Mathematical Models of Solid with Rheological Properties

Fig. E.12 Five-parametric structural model and its graph of strain mode

14. A cylinder of viscoelastic material is tightly placed in a rigid container such


that ε11 = ε22 = εrr = 0 (no radial strains), (Fig. E.13). The body show elastic
properties at dilatancy and creep in accordance with the law Js = J0 (1 + t),
where J0 is constant. Determine σ33 (t), if ε̇33 = A = constant.
Answer: [ ]
σ33 (t) = A K t+4(1−exp(−t))
3J0
U (t).
15. A viscoelastic body in the form of a block shows elastic properties at dilatancy
and viscoelastic properties at shear in the form of Maxwell’s law. The unit is
subjected to the pressure σ11 (t) = p0 δ(t) distributed uniformly over surface x 1 ,
(Fig. E.14. If the block strains are constrained so that ε22 = ε33 = 0, define
ε11 (t) and σ22 (t).

Fig. E.13 Principal scheme


for exercise 14

Fig. E.14 Principal scheme


for exercise 15
6.4 Equations of Mechanics of Linear-Hereditary Media 147

Answer: ( )
σ22 (t) = p0 (2G−3K )δ(t)
− 6G
exp([−3K /((3K + 4G)τ)]t) U (t),
( 3K +4G 3K +4G
)
ε11 (t) = p0 3K3δ(t)
+4G
− (3K +4G)K
4G
exp([−3K /((3K + 4G)τ)]t) U (t).
Chapter 7
Mathematical Models of Plasticity
Theory

7.1 Solid-Plastic Behavior Models

Plasticity is the ability of a solid to deform under the external loads and to obtain
residual (plastic) strains after load removal. At the same time, there is no one-to-one
relationship between stresses and strains occurring in the body. That is, it is impos-
sible to find strains from stresses and, inversely, knowing strains cannot determine
stresses.
Plastic strains before destruction reach the values of 10–20%, while elastic
ones reach only 0.3–0.5%. Therefore, strength calculations only based on the
permissibility of elastic strains are often not technically and economically feasible.
Recall that there is a fundamental difference among elastic, plastic, and
rheological models of medium behavior.
• After changing the load of the elastic medium, its state is uniquely determined by
the new conditions and does not depend on the previous one.
• Plasticity models “remember” the history of previous loads. However, since the
change of states occurs instantly, in plasticity models this story is not associated
with time.
• If the rock’s plasticity characterizes their behavior at stresses greater than the
elasticity breaking point, then creep, which is a slow increase in plastic deforma-
tions, is also exerted at stresses less than the elasticity breaking point, but with a
sufficiently long-term loading action.
Material behavior completely different from viscoelasticity can be described using
the Coulomb element (“dry friction” element) (Fig. 7.1).
In this case, the strain rate is zero until the absolute stress value reaches the yield
stress σ y .

|σ| < σ y ⇒ ε̇ = 0. (7.1)

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 149
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_7
150 7 Mathematical Models of Plasticity Theory

Fig. 7.1 Coulomb element


representation and its
behavior diagram

Yielding deformations can reach any value in the direction of the applied force
after reaching yield stress, that is:

σ = +σy ⇒ ε ≥ 0 arbitrary,
σ = −σy ⇒ ε ≤ 0 arbitrary.

During the process of plastic yielding, the internal energy of the solid dissipates
regardless of the strain rate:

πi = σε̇ = σ y |ε̇| = δ ≥ 0.

This material behavior is called the rigid-perfect plasticity. In this case, there
is no functional relationship between current stresses and strains. In addition, this
condition is independent of the strain rate (rate-independent) (i.e., scleronomic).

7.2 Material Plasticity During Tension and Compression

The elastic and plastic properties of materials are showed with a rather slow, so-called
static or quasi-static action of external loads. In this case, the deformation properties
of materials practically do not depend on the time, the rate of loading increase and
the duration of external loading action.
The basic plastic properties are conveniently considered in the simplest example
of tension–compression of a cylindrical sample. We consider a metal sample, which
before the experiment beginning is isotropic and has the same yield strength for
tension and compression. We will denote the tensile stress in the sample through σ
and the relative elongation through ε.
7.2 Material Plasticity During Tension and Compression 151

Fig. 7.2 Typical diagrams σ~ε the sample tensile

As it is known, the tensile dependence of the sample σ~ε for most materials has
a straight start section until the stress exceeds the breaking point σpr . In this area,
the Hooke’s Law is valid. The curved section follows the straight section. The linear
connection between stresses and strains is broken in this area. At the same time,
some materials have a clearly defined plastic yielding area (Fig. 7.2a), and some do
not (Fig. 7.2b). In this case, a conditional yield stress is accepted: stress at which
the plastic strains, remaining after the load removal, is 0.2% or 0.002. It is indicated
by σ0,2 .
The point A of the tensile diagram of the sample (see Fig. 7.2a) corresponds to
the stress σ and the total strain ε, which, in turn, can be divided into elastic strain
εe and plastic strain εp . After unloading, the elastic strain disappears and the plastic
strain remains.
When the sample is being unloaded, the relationship between stresses and strains is
linear with the slope of the corresponding line being the same as the original section.
Thus, when the sample is unloaded, the sample behaves as an elastic material.
If the sample is stretched again after full unloading, as shown in Fig. 7.2c, the linear
section will continue to greater values σ = σ'y than it was in the first time (σ'y > σ y ).
Therefore, when the sample is re-loaded after plastic deformation, the proportional
breaking point of the material increases. This phenomenon is called hardening.
Thus, for real materials, the yield stress is not constant, but develops during plastic
deformations. In general, the yield stress may increase due to the deformation (this
is hardening as previously stated), but the yield stress may decrease. This behavior
is called softening.
For many materials, hardening is first observed to some deformations and then
softening occurs to the moment of destruction. When the sample is being softened, the
material becomes unstable and disposed to localization like shear strips or the appear-
ance of a “yield neck” (Fig. 7.3). Therefore, in the softening mode, the determination
of the relations between stresses and strains is rather complicated.
Determining the ultimate stiffness also becomes difficult. Therefore, it is usually
limited to consideration of the hardening regime. In this case, the following hardening
rules can be adopted:
• Linear hardening law σ y = K |ε| + σ y0 , K is positive constant;
152 7 Mathematical Models of Plasticity Theory

Fig. 7.3 Hardening and


softening phenomena of the
material

• Or potential hardening law σ y = K |ε|n + σ y0 , K, n are positive constants.


Thus, when the material is re-loaded from tension to compression or vice versa,
there may be a difference between the yield stresses in both modes. That is, the
material enters plastic area under repeated loading at other stresses than the initial
limit of proportionality σ y . Usually, when the sample is loaded with the load of an
opposite sign (compression), the material will enter the plastic area at stresses less
than the initial limit of proportionality, i.e., σ''y < σ y (Fig. 7.2c).
Such a phenomenon was studied in detail by I. Bauschinger and is called by
his name—the Bauschinger1 effect, (Fig. 7.4). If equality is achieved (as far as the
tensile yield stress is increased, by the same amount it decreases under compression),
then the material is called cyclical ideal. If the equality σ'y + σ''y = 2σ y is true
(how much the tensile yield stress increases, by the same value it decreases under
compression), then the material is called cyclically ideal.
Thus, the pre-stretched sample has an increased tensile proportional limit and a
reduced compression limit due to hardening.
Two types of hardening are described in the literature. The difference between
them is observed when the load is reversed, (Fig. 7.5) [Bertram, A. Solid Mechanics.
Theory, Modeling, and Problems/Albrecht Bertram, Rainer Glüge. –Springer Inter-
national Publishing Switzerland, 2015. 331 p.]:
1. Isotropic hardening, for which the yield limit under tension and compression
changes in parallel;
2. Kinematic hardening, for which the yield limit under tension increases in a way
other than its reduction under compression.

1 Johann Bauschinger (1834–1893), mathematician, builder, and professor of Engineering


Mechanics at Munich Polytechnic from 1868 until his death. The Bauschinger effect in materials
science is named after him.
7.2 Material Plasticity During Tension and Compression 153

Fig. 7.4 Saturation hardening (a) and Bauschinger effect (b)

Fig. 7.5 Two types of hardening

The softening behavior often lies between these two extremes, which can be
described using a combination of the two hardening models. For that purpose, we
need two hardening variables, namely one for the isotropic hardening σ y and a back
stress σb for kinematic hardening. For both variables, the evolution equations are
needed.
As the experiments show, a certain effect on the deformation curve of the sample
is the rate of loading increase or the strain rate. In conventional test machines, the
strain rate of the samples varies within ε̇ = 10−5 − 10−2 c−1 . This deformation mode
is called static. Static test diagram is independent of strain rate. This dependence is
noticeably manifested, starting with order rate of 10−1 c−1 . Schematic diagrams of
small-carbon steel tests obtained at three different levels of strain rates are shown
in Fig. 7.6: curve 1 corresponds to the value ε̇ = 10−4 c−1 ; curve 2 corresponds to
ε̇ = 0.5 c−1 , and curve 3 corresponds to ε̇ = 102 c−1 .
These comparisons show that the modulus of elasticity at uniaxial tension is
practically unchanged. Stress and strain yield limits increase with increasing tension
rates. Such expansion of elastic deformation range is related to the inertia of plastic
deformation mechanism.
154 7 Mathematical Models of Plasticity Theory

Fig. 7.6 Deformation


modes at different strain rates

An analytical representation of the dependence of the yield limit value on the


strain rate can be taken as linear:
·
σ y = σ y0 + μ0 ε,

where σy0 is the static elasticity limit; μ0 is the experimentally determined constant
of material.
With an increase of the deformation rate, the material ultimate tensile stress σb
(stress corresponding to the fracture) increases, and the corresponding strain (ultimate
plastic elongation) decreases. This phenomenon is called material embrittlement.
Brittle material is destroyed without noticeable plastic deformations preceding the
destruction.

Remark It should be noted that there are still no reliable systematic experimental
studies on the effect of strain rates for many materials. This is due to the fact, that
the limit rates at which the described effects are observed are not widely used in
practice. Therefore, in most cases, the static diagram of the tensile test material is
sufficient to characterize its mechanical properties.
A common approach is that the plastic properties of the material are characterized
by a plastic coefficient. The plastic coefficient is the ratio of the work W p spent on
destroying a given volume of real material to the work W i.e required to destroy the
same volume of material, with the same value of compressive strength under the
assumption of material ideal elasticity:

K pl = W p /Wel .
7.3 Elastic–Plastic Models of Material Behavior 155

Fig. 7.7 Prandtl or Genky


structural model

7.3 Elastic–Plastic Models of Material Behavior

Therefore, the behavior of many real materials is first elastic until the deformations
reach plastic yield limit. Only upon reaching the yield limit σ y (often referred to as
the elastic limit), the material deforms plastically (plastic yielding).
This behavior can be modeled by placing in a series of Coulomb and Hooke’s
elements. The constructed model is called the Prandtl2 model or the Genky3 model
(Fig. 7.7).
Consider a one-dimensional case.
The total deformation in the Prandtl model can be represented as the sum of the
elastic and plastic components:

ε = εl + ε p , εl = σ/E. (7.2)

In the hardening mode, only plastic deformations appear ε p .


Stress power for the Prandtl model is:
( ) | |
πi = σε̇ = σ ε̇l + ε̇ p = Eεl ε̇l + σ y |ε̇ p | = ẇ + δ.

That is, in this case, the stress power consists of elastic and dissipative parts.
The elastic–plastic model with kinematic hardening is, for example, the Masing
model (Fig. 7.8).
For metals, the elastic regions in the strain space are often small enough to
justifiably use only the laws of the linear theory of elasticity.

2 Ludwig Prandtl (1875–1953), German fluid dynamitist, physicist, and aerospace scientist.
3 Heinrich Hencky (1885–1951), German engineer-mechanics.
156 7 Mathematical Models of Plasticity Theory

Fig. 7.8 Masing structural model

If one considers elastoplastic material behavior under monotonous loading


without loading reversal, one cannot distinguish its behavior from a non-linear
elastic behavior. This fact inspired Hencky to suggest a finite deformation rule,
which allows for treating the material behavior in the context of non-linear elasticity
as long as no loading reversal occurs. A similar approach is the Ramberg–Osgood
law:
( )N
σ σ
ε= + k2 , k1 , k2 , k3 , N − constants.
k1 k3

Plasticity theories establish a connection between stresses and strains


(deformation theories) or strains rates (plastic flow theory) in the areas of the
plasticity of materials. At the same time, stresses often depend not only on current
strains, but also on the history (process) of deformation.
The deformation theory of plasticity has a more limited area of application than
the plastic flow theory, since it has a wider range of restrictions.
7.5 Plasticity Conditions 157

Fig. 7.9 Viscoplastic structural model

7.4 Viscoplasticity Models

In addition to non-linearity, speed independence is a characteristic of all elasto-


plastic models, which exclude creep and relaxation. Viscoplasticity models combine
plasticity and creep.
A simple method to include viscous effects is adding viscous elements like
dampers in series (V 1 ) or in parallel (V 2 ) (or both) to the plastic elements (Fig. 7.9)
[Bertram, A. Solid Mechanics. Theory, Modeling, and Problems / Albrecht Ber-tram,
Rainer Glüge. –Springer International Publishing Switzerland, 2015. 331 p.].
The effect of V 1 is that there are no more elastic ranges but rather viscoelastic ones.
Under deformation cycles even below the yield limit, one will observe a hysteresis
(be similar to the Maxwell model).
What is even more important is the application of a viscous element V 2 in
parallel with the Coulomb model. V 2 is only activated under inelastic deformations.
Therefore, elastic ranges may still exist.
During yielding, the stresses are composed by the yield stress σ y and a viscous
overstress σ0 : σ = σ y + σ0 .
As a consequence, the stresses can be arbitrarily large or small, which is perhaps
more realistic than the restriction to ±σ y for a perfect plastic material.
If one combines such models with complex hardening mechanisms, one can
describe rather complex material behavior. Common of all such suggestions is the
intention to describe the material behavior under essentially all kinds of loadings
(creep, cyclic loads with or without holding time, relaxation, etc.) as realistically as
possible.

7.5 Plasticity Conditions

For the purpose of simplification, in the general case, non-linear relations between
stresses and strains, the dependencies σ~ε for real materials are often approximated
in the form of piece-broken lines (Fig. 7.10a–c).
158 7 Mathematical Models of Plasticity Theory

Fig. 7.10 Various approximations of the dependencies σ~ε for real materials

The simplest is the Prandtl diagram for a perfectly elastoplastic material,


(Fig. 7.10a). A diagram with linear hardening is shown in Fig. 7.10c. These two
approximations are most often used in solving problems of plasticity theory.
The tension and compression curves “stress–strain” for most materials are very
close, and we will consider them to be coincidental in the future.

Remark More careful experiments show that the unloading law is not always linear.
In existing plasticity theories, these minor deviations from Hooke’s law during
unloading are neglected, as the difference between the proportional limit and the
yield limit.
Consider a body of arbitrary shape, and consider that there are no initial stresses
and strains in it. At the initial stage of such solid loading, only elastic deformations
occur and, therefore, the appearance of plastic deformations is uniquely determined
by the current stresses. Therefore, the plasticity condition can be written as some
functions of the stress tensor components.
For an isotropic material, the plasticity condition cannot depend on the choice
of coordinate system. Therefore, the corresponding function must be determined by
three stress tensor invariants, which can be, for example, three principal stresses:

f (σ1 ,σ2 ,σ3 , k) = 0, (7.3)

where k is a constant value, characteristic of material (for example, yield limit σ y ).


For example, with uniaxial tension, we have: σ1 = σ y .
As experiments show, plastic deformations are associated with shear
phenomena. Therefore, the basic plasticity criteria, which are widely used, compare
some tangent stresses with yield stress limits. Therefore, according to the Tresca-
Saint-Venan condition, the transition from an elastic state to a plastic state occurs
if the maximal tangent stress reaches a certain limit value for a given material:

|σ1 − σ3 | σy
|τmax | = = or σ1 − σ3 = σ y , (7.4)
2 2
where, as before, the principal stresses are numbered in decreasing order: σ1 ≥ σ2 ≥
σ3 ; σ y is the tensile yield limit. Solid is assumed to be initially isotropic.
7.5 Plasticity Conditions 159

Remark In the course of mechanics of materials, the Tresca-Saint-Venan criterion


is used as the criterion of maximal tangent stresses: σred = σ1 − σ3 ≤ [σ].

Another plasticity criterion determines the transition from an elastic state to a


plastic state of an initially isotropic material, if:

σu = σ y , (7.5)

where σu is the stress intensity (see Chap. 2). σu is proportional to the octahedral
tangent stress and its square is proportional to the second invariant of the stress
deviation:
( )
σu2 = −3J2d = (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 /2.

In this case, not the maximum, but the octahedral tangent stress reaches the limit
value for this material. This criterion is used in the energy theory of strength and the
name is the Mises (Huber4 —Mises—Genky) plasticity condition.
Results obtained by plasticity criteria (7.4), (7.5) are very close. When solving
problems, it is usually recommended to use the criterion that simplifies the solution.
Since the solid deformed substantially and elastically during all-round tension or
compression, it can therefore be assumed that the plasticity condition generally does
not define all the stress tensor, but only its deviatoric part.
As we have already said, the transition to the plastic state cannot depend on
the choice of the coordinate system, so the plasticity condition is a function of the
invariants of the stress deviator. The first invariant is zero (J 1d = 0), so in the general
case the condition for the appearance of plastic deformations is determined by the
second and third invariants of the stress deviator:

f 1 (J2d , J3d ) = 0. (7.6)

This equation in the coordinate system σ1 , σ2 , σ3 describes a surface that is called


the yield surface.
Condition (7.6) contains the invariants of the stress deviator and some material
constants, e.g., yield limit.
These plasticity criteria make it possible to fix the moment of the first appearance
of plastic deformations.
These criteria are sufficient to solve plastic problems when the material deforma-
tion in a uniaxial stress state respected to the Prandtl diagram (Fig. 7.10a). This is
explained by the fact that during such materials repeated loading there is no change
in the condition of plastic deformations.
The situation changes if the material is hardened (Fig. 7.10b, c). Such re-loading
materials are characterized by an increase in yield strength, the value of which

4Huber Maximilian Tytus (1872–1950), a mechanical scientist, was president of the Academy of
Technical Sciences of Poland.
160 7 Mathematical Models of Plasticity Theory

depends on the accumulated plastic deformation. In such cases, a hardening condition


that looks like a plasticity condition, (7.6) should be introduced:

f 1 (J2d , J3d ) = .(η). (7.7)

Condition (7.7) includes a function .(η) that depends on the material hardening
parameter η. This equation in the space of principal stresses also determines the yield
surface, the change in position, shape, and size of which during loading characterizes
the material deformation hardening.
For example, if the hardening parameter coincides with the strain intensity, then
one of the variants of the criterion (7.7) can be taken as a generalization of the Mises
condition:

σu = .(εu ), (7.8)

which are examples of the yield surface (7.6), construction for the considered
Tresca-Saint-Venan and Mises plasticity criteria.
In the case of the Mises criterion (7.5), the yield surface equation (σu = σ y ) can
be written as:

(σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 − 2σ2y = 0.

This equation in axes σ1 , σ2 , σ3 describes a cylinder, the axis of which is equally


inclined to coordinate axes (Fig. 7.11). If you cut the cylinder with a plane σ3 = 0,
the section will be an ellipse whose equation is:

σ12 − σ1 σ2 + σ22 = σ2y .

Fig. 7.11 Yield surface for


Mises criterion
7.5 Plasticity Conditions 161

Fig. 7.12 Yield surface for Tresca-Saint-Venan criterion

Therefore, the Mises plasticity criterion corresponds to the yield surface in the
form of a circular cylinder, the
√ radius of which in the plane perpendicular to the axis
of the cylinder equals to σ y / 2.
If you accept the Tresca-Saint-Venan criterion (σ1 −σ3 = σ y ), taking into account
the fact that the number of the principal stresses is not always known in advance,
then the following equalities may occur:

σ1 − σ3 = ±σ y , σ3 − σ2 = ±σ y , σ2 − σ1 = ±σ y .

The yield surface in this case is represented as a hexagonal prism with an axis also
equally inclined to the axes σ1 , σ2 , σ3 (Fig. 7.12a). This prism is called the Coulomb
prism. It is inscribed in the Mises cylinder. Prism and cylinder axes coincide. The
equation of this axis is σ1 = σ2 = σ3 . Figure 7.12b shows the cross section of the
cylinder and prism with a plane corresponding to the plane stress state.

Remark In order to take advantage of the Saint-Venan plasticity condition, it is


necessary to know in advance which of the principal stresses is maximal and which
is minimal. When using the Mises condition, you do not need to define the principal
stresses at all.
The condition σu = σ y can be written using the definition of the stress intensity:

(σ11 − σ22 )2 + (σ22 − σ33 )2 + (σ33 − σ11 )2 + 6(σ12


2
+ σ23
2
+ σ31
2
) = 2σ2y .

Therefore, in many cases, the use of the Mises plasticity condition is more conve-
nient than the Saint-Venan conditions. The maximal difference in calculations for
these criteria does not exceed 13% (case of pure shearing).
162 7 Mathematical Models of Plasticity Theory

7.6 Simple and Complex Loading

In defining relations of plasticity theory (the relationships between stresses and


strains), obviously, not only the current values of the stress and strain tensor
components should be taken into account, but also the ways in which they are
achieved.
In the theory of plasticity, two types of solid loadings are distinguished: simple
and complex loadings.
Recall (see Chap. 2) that loading is called simple if all stress tensor components
increase in proportion to one common parameter λ (for example, time), that is σi j =
λ σi0j . In this case, the components of the guide tensor (2.28) remain unchanged.
Otherwise, loading is called complex.
Examples of simple and complex loadingof simple and complex loading.
Assume that the cylindrical tube is under uniform axial stretching and torsion
(Fig. 7.13). If the tube has a sufficiently thin wall, the stress state therein can be
considered as plane. Normal σx and tangent τ stresses in the tube are from known
expressions:

P mz
σx = , τ= . (7.9)
2πRδ πR 2 δ
When the values of external loads P and mz change in proportion to one parameter
λ, for example, time, simple loading is carried out, since the of the stress tensor
components, according to (7.9), change in proportion to the same parameter. The
loading trajectory is in the stress axes σx , τ is shown in Fig. 7.14.

Fig. 7.13 Example of


simple loading

Fig. 7.14 Diagrams of


various loads
7.6 Simple and Complex Loading 163

Simple loading corresponds to beam OA. Thus, for simple loading, the following
equation is true:

σi j = λ σi0j ,

where σi0j are some initial values of stress tensors at the loading application.
Then the average stress σ, the stress intensity σu and the modulus of deviator stress
s, due to formulas (2.27), (2.28), will also be linear with respect to the parameter λ:
Σ
σ = λ σ0 , σu = λ σu0 , s = 3/2σu = λ s 0 .

In turn, the components of the stress guide tensor (2.28) do not depend on λ, since:

σi j − σδi j λσi0j − λσ0 δi j


si j = = = s i0j .
s λs 0
Now let’s consider an example with another load (Fig. 7.14), in which an axial
load P was first applied to the tube, creating a normal stress. After the value of the
normal stress has reached the value σ∗ , the torque mz was applied. The normal stress
σ∗ during torque application remains unchanged and the tangent stress increases from
zero to a value τ∗ . As a result, we again get point A, but along a broken trajectory.
Such loading is complex. A similar result would be achieved by first twisting the
tube to a value τ∗ , and then stretching it to a stress σ∗ .
Note that for an elastic body, the sequence of its loading does not play a role, since
there is a one-to-one relationship between the stress and strain states, regardless of
how they are created. In elastoplastic bodies, the situation turns out to be funda-
mentally different. In this case, not only the nature of the stress state at its points
is significant, but also the path along which it was created. Depending on this, the
deformed state at the same points of the body can change significantly.
Prof. A.A. Ilyushin proved the theorem on sufficient conditions of simple loading.
This requires the proportionality of the external load to one common parameter and
the power dependence of stress intensity on strain intensity:

σu = A εαu ,

where A and α are constants. At α = 0, this equation coincides with the Mises
plasticity condition σu = const. At small values of α, it gives curves with low
hardening dσu /dεu . If α = 1 we have Hooke’s law. Note that this condition is
sufficient, but not necessary. Simple loading may occur in some cases where the
above condition is violated.
Active and passive deformation processes are distinguished in studies of the
plasticity.
164 7 Mathematical Models of Plasticity Theory

With an active process, also called a loading increment of work performed by


external loads over the body is positive. Negative increment of this work corresponds
to passive process, or unloading.
For elementary material volume, the active process condition can be written as
the following inequality:

σi j δεi j > 0,

and passive process condition as follows:

σi j δεi j < 0.

Sometimes an active process is taken as such that the plastic deformation


increases, and a passive process is taken as such that the plastic deformation remains
unchanged.
In plasticity theories, it is assumed that at any moment of a deformation process,
the tensor of complete strain is represented as the sum of elastic and plastic strain
tensors:
p
εi j = εiej + εi j . (7.10)

Moreover, the first tensor changes in both active and passive processes, while the
plastic strain tensor changes only in the active process.
Plastic deformation is defined as the collection of strain tensor components that
remain at the solid point when all stress tensor components at that point turn to zero.
Thus, plastic deformations are identified with residual deformations.

7.7 Hypotheses of the Small Elastoplastic Deformation


Theory

The theory of small elastoplastic deformations is the simplest theory of a hardening


plastic material. It is very common in applications. This theory refers to deformation
theories of plasticity.
Consider the slow rod extension process, (Fig. 7.15).
A rod is loaded along the section OAB, a line BC corresponds to unloading. The AB
loading branch equation σ1 = f (ε1 ) can represent both plastic and non-linear elastic
deformation of the rod. In connection with this remark, it is possible to attempt
to construct plastic deformation equations in the form of finite relations between
stresses and strains. Such equations would be significantly simpler than the theory
of plastic flow equations. Therefore, in the theory of small elastoplastic deformations,
it is assumed that for elastoplastic bodies, it is possible to establish relationships
between stresses and strains, like Hooke’s law for elastic bodies.
7.7 Hypotheses of the Small Elastoplastic Deformation Theory 165

Fig. 7.15 Rod


extension-compression
diagram

А В

О С

The theory of small elastoplastic deformations is based on the hypotheses


proposed by M.T. Huber, R. Mises, G. Genki and generalized in the case of a mate-
rial with hardening by A. Nadai. The development and substantiation of the theory
of small elastoplastic deformations was carried out in the works of A.A. Ilyushin.
Therefore, often the theory of small elastoplastic deformations is called the Ilyushin
plasticity theory.
In the theory of small elastoplastic deformations, it is assumed that with a simple
active deformation of an initially isotropic material, the properties of which do not
depend on the third invariant of the stress tensor, the following three hypotheses are
true.
1. Volume strain is elastic. Therefore, the volumetric strain of the body θ is consid-
ered elastic. It is directly proportional to the average normal stress σ and for it the
Hooke’s law is valid: σ = K θ = 3K ε. This means that the volume changes only
due to elastic deformations, and the material behaves as incompressible materials
during plastic deformation. Thus, due to plastic deformation, the body volume
does not change.
2. Coaxiality of stress and strain deviators. Strain deviator components .ij are
proportional to stress deviator components sij . The relationship between them
can be represented in the form proposed by A.A. Ilyushin:

2 σu
si j = .i j , (7.11)
3 eu

where σu , εu are intensities of the stress and strain tensors.


In stress and strain components, the expression (7.11) is:

2 σu
σi j − σδi j = (εi j − eδi j ), (i, j = 1, 2, 3). (7.12)
3 eu
( ) ( )
where σ = σx x + σ yy + σzz /3; e = ex x + e yy + ezz /3;
166 7 Mathematical Models of Plasticity Theory

√ /( )2 ( )2 ( )
σu = 22 σx x − σ yy + σ yy − σzz + (σzz − σx x )2 + 6 σx2y + σ yz 2 + σ 2 is
zx
the stress intensity;
√ /( )2 ( )2 ( )
eu = 32 ex x − e yy + e yy − ezz + (ezz − ex x )2 + 6 ex2 y + e2yz + ezx 2 is the

strain intensity.
Similar to elastic deformation in the model of small elastoplastic deformations in
the area of plastic deformations, the values of stresses and strains are connected to
each other by unique dependence.
It has been shown experimentally that the law of the relationship between stresses
and strains (7.12) is fulfilled for the case of proportional or simple loading.
3. Hardening hypothesis. Regardless of the type of stress state for each material,
there is a universal relationship between stress intensity and strain intensity:

σu = .(εu ). (7.13)

Remark For an elastic material, this relationship is expressed by a linear relationship:


σu = 3G εu .
Experimental tests of these hypotheses gave good enough results for simple
loading or little different from simple loading. The deformation process must be
active without unloading.

Remark The theory of small elastic–plastic deformations was proposed by Genki


and is based on finite dependencies between stress components and strain compo-
nents. Genki equations are written in the following form:
( ) ( )
εi j − δi j ε0 = . σi j − δi j σ0 , (7.14)

where ε0 = σ0 /3K , K = E/(1 − 2ν).

The function . is defined as follows:


3 εi
.= , (7.15)
2 σi

where εi is the strain intensity; σi is the stress intensity.


Then the Eq. (7.14) can be represented as:

( ) 3 εi ( )
εi j − δi j ε0 = σi j − δi j σ0 . (7.16)
2 σi

The inverse dependencies between the stress deviator components and the strain
deviator components are recorded in this way:

( ) 2 σi ( )
σi j − δi j σ0 = εi j − δi j ε0 . (7.17)
3 εi
7.7 Hypotheses of the Small Elastoplastic Deformation Theory 167

As can be seen, the representation (7.17) is similar to (7.12).


There is a certain relationship between stress intensity and strain intensity
σi = .(εi ), with small elastoplastic deformations for each material, similar to
the relationship between stress and strain in tension σ = f (ε). The relationship
σi = .(εi ) can be plotted from the tension pattern. To do this, you must first replace
σi with σ, and εi with (ε − ε0 ).
Since the theory of small elastoplastic deformations assumes that the volume
does not change during plastic deformations (θ = 0), it is necessary to put ε0 = 0
in Eqs. (7.16) and (7.17). In this case, the relationship between stress and strain
intensities σi = .(εi ) beyond elasticity is determined by the tension diagram. Thus,
the hypothesis is accepted that there is a “single deformation curve” for a material,
regardless of the stress state type. Under conditions of material incompressibility:
K → ∞; μ = 0.5; E = 3 μ.
One of the varieties of the deformation theories of Genka, characterizing the
relationships between the components of strains and stresses for elastoplastic
deformations, has the form:
( )
1+ϕ ϕ + 1+ν

εi j = σi j − δi j σ0 . (7.18)
2μ 1+ϕ

In these relationships, the value of shear modulus is reduced by 1/(1 + ϕ), i.e.,
the material becomes less rigid.
The inverse dependencies between stress and strain components are recorded as
follows:
( )
2μ ϕ + 1+ν

σi j = εi j + δi j ε .
3ν 0
(7.19)
1+ϕ 1 − 1+ν

If in Eq. ((7.19) we express


) strains through displacements along Cauchy formulas
(i.e., εi j =
Σ u i, j + u j,i /2, and then we substitute these formulas into equilibrium
equation σi j, j + X i = 0, then we get three differential equations relative to four
j
unknown functions (u i , ϕ). We add to these three differential equations a fourth,
representing the Huber-Mises-Genky plasticity condition:
( )2 ( )2 ( )
σx − σ y + σ y − σz + (σz − σx )2 + 6 τ2x y + τ2yz + τ2zx = 2σ2y .

In this equation, we replace stresses with representations through displacements


u i and the function ϕ.
As a result, we obtain four equations to determine four unknown coordinate
functions.
From the relationships (7.16) and (7.17), it follows the proportionality of the stress
deviator components to the strain deviator components, as well as the proportionality
of the principal angular strains to the principal tangent stresses.
168 7 Mathematical Models of Plasticity Theory

A.A. Ilyushin on the basis of experimental data showed that Genka equations are
confirmed for simple loading processes or loading processes close to simple ones.
The equations of deformation theory of plasticity are the equations of a nonlinear
elastic solid. Therefore, using these equations to describe plastic deformations in
complex loading paths can lead to unsatisfactory results. Therefore, we can assume
that the equations of the deformation theory of plasticity are suitable if the plastic
deformations develop in a certain direction.

7.8 Theory of Problems of Small Elastic–Plastic


Deformations

We consider an elastic–plastic body that is under the influence of mass forces ρF i and
surface loads Ri . For the solution of the theory of problems of small elastic–plastic
deformations (that is, for definition of unknown displacements, strains, and stresses
(ui , σij , εij ; i, j = 1, 2, 3)), there are equilibrium equations, Cauchy ratios, strain
compatibility equations, and boundary conditions:

σi j , j + ρ Fi = 0; εi j = (u i , j +u j ,i )/2;
εαα ,ββ + εββ ,αα − 2εαβ ,αβ = 0,
(7.20)
(εαβ ,γ − εβγ ,α + εγ α ,β ),α − εαα ,βγ = 0;
u i = u i0 (x) on Su , σi j l j = Ri on Sσ .

Constitutive equations are conveniently written as:

si j = 2G ϕ(εu ) .i j , σ = 3K ε, (7.21)

where si j = σi j − σ δi j ; .i j = εi j − εi j δi j ; i, j = 1, 2, 3.
Comparing the dependencies (7.12) and (7.21), we can find the expression of the
plastic function ϕ(εu ) through the previously introduced universal function F(εu )
(7.13):

.(εu )
ϕ(εu ) ≡ .
3Gεu

Equations (7.21) are valid only in the loading process. In the case of elastic
unloading from the generalized Hooke’s law, such ratios follow:

σi j − σi' j = 2μ(εi j − εi' j ) + λ(θ − θ' )δi j , (7.22)

where σi' j and εi' j are stresses and strains existing before the start of unloading; λ
and μ are Lamé constants.
7.9 The Method of Elasticity Solutions 169

Unloading equations in form (7.22) are maintained until new (secondary) plastic
deformations appear during unloading.
When we solve elastic–plastic problems, the obtained solution must satisfy not
only forces and kinematic boundary conditions, but also additional conditions on
the interface of the zones of elastic and plastic deformations.
The problem of plasticity theory is non-linear, so the question arises with existence
and uniqueness of the solution. A.A. Ilyushin ([Ilushin, A.A. Plasticity. Fundamentals
of general mathematical theory / A.A. Ilushin. – Moscow. Pub. by Academy of
Sciences USSR, 1963. 272 p. (in Russian)]) proved that the system of Eqs. (7.20),
(7.21) is of the elliptical type, if the following condition is met:

σu d σu
3G ≥ ≥ > 0. (7.23)
εu d εu

At the same time, a solution to this problem exists if there is a solution to the
corresponding problem of the linear theory of elasticity.
A.A. Ilyushin proved the following theorem regarding the uniqueness of the solu-
tion: “With given body forces ρF i and surface forces Ri on a part of the boundary
surface S σ and displacements ui on a part of the boundary surface S u , the stress–strain
state of the body (that is, ui , σij , εij ) are determined in the only way if the loading is
simple”.

7.9 The Method of Elasticity Solutions

The solution of the problems of plasticity theory is associated with the solution
of a system of non-linear differential equations in partial derivatives, which is an
extremely complex mathematical problem, which in analytical form is solved, as a
rule, in exceptional cases. Therefore, most often used approximate methods. One of
them is the method of successive approximations proposed by A.A. Ilyushin to
solve the problems of the theory of small elastic–plastic deformations with active
loading and called in the theory of plasticity the method of elasticity solutions. The
essence of this method is to consider the sequence of linear problems of the theory
of elasticity, the solutions of which with an increase in ordinal number converge to
the solution of the problem of plasticity theory.

ϕ(εu ) = 1 − ω(εu ). (7.24)

The representation (7.24) is substituted into the expression (7.21), then the
physical relations become as follows:

si j = 2G(1 − ω(εu )).ij , σ = 3K ε, (7.25)

where 0 ≤ ω < 1; ω(εu ) = 0, if εu ≤ ε y .


170 7 Mathematical Models of Plasticity Theory

Thus, with ω = 0, Eq. (7.25) coincide with the relations of the linear theory of
elasticity.
We write the differential equilibrium equation and boundary conditions by
decomposing the stress tensor into the deviatoric and spherical parts:

si j , j +σ,i +ρ Fi = 0,
(7.26)
si j l j + σli = Ri on Sσ , u i = u i0 (x) on Su .

We solve the problem of elastic-plasticity in displacements. Therefore, compo-


nents (7.25) are substituted into equilibrium equations and force boundary conditions
(7.26) and Cauchy ratios are taken into account. As a result, generalizations of Lamé
equations are obtained:

(λ + μ)θ,i +μ .u i + ρ Fi − ρ Fωi = 0, (7.27)

λθ li + μ(u i , j +u j ,i )l j = Ri + Rωi ,

where Fωi = 2G(ω .ij ),j ; Rω i = 2G .i j l j ω .


For zero approximation ω(0) = 0 is taken. Then, fictitious loads are Fωi = Rωi =
0. Therefore, to determine the first approximation u i(1) , we have the usual problem
of the linear theory of elasticity. The following components are defined from the
determined displacements u i(1) :

.i(1) (1)
j , εu , ω
(1) (1)
≡ ω(εu(1) ), Fωi , Rω(1)i .

For any kth approximation, there are such equilibrium equations and boundary
conditions:

(λ + μ)θ,i(k) + μ.u i(k) + ρ Fi − ρ Fωi


(k−1)
= 0,
λθ(k)li + μ(u i(k) , j −u (k) (k−1)
j ,i )l j = Ri + Rωi , (7.28)

where Fω(k−1)
i , Rω(k−1)
i determined by the preceding (k-1)th approximation.
Modified Lamé equations and boundary conditions (7.28) are linear with respect
to unknown displacements u i(k) . They differ from the corresponding equations of the
theory of elasticity in those fictitious forces Fω(k−1)
i , Rω(k−1)
i are added to the external
forces ρF i , Ri . This allows, according to the known solution of the corresponding
problem of the theory of elasticity, to construct a solution of the elastic–plastic
problem in a recurved form.
Remark An example of building such an iterative solution is discussed in the
Sect. 7.12.
For convergence of the methods of elastic solutions, it is necessary that the param-
eter ω associated the relation (7.24) with the function ϕ(εu ) is small compared to the
unit. At the same time, the following condition must be met:
7.9 The Method of Elasticity Solutions 171


1 > ω + εu ≥ ω ≥ 0.
dεu

Convergence of this method has been investigated by various authors. In practice,


it has been found that the convergence rate of the methods of elastic solutions is very
high, so that several approximations are enough to obtain the necessary accuracy.
Let’s give an algorithm for solving the problems of elastic-plasticity by the
methods of successive approximations, namely, by the method of elastic solutions.
There are several ways to implement this method. Consider two of them: the
method of additional loads and the method of additional deformations. Using them
allows you to solve elastic–plastic problems using the approaches of the theory of
elasticity to determine functions u i (x) and σ i j (x).

Remark Note that we are talking about the theory of plasticity problems without
unloading zones when using the deformation theory of plasticity.
Let’s consider an approach based on the method of additional loads.

Equilibrium equations obtained by the method of additional loads have the


following form:
( ) ( )
μ.u j + (λ + μ) ∂ θ/∂ x j + X j + X 0j = 0, j = 1, m , (7.29)

where
[( )( )] Σ m [( )( )]
∂ 1 ∂u j 1 ∂ 1 ∂u j ∂u k
X 0j =− 2μ − − θ − μ− + .
∂x j ψ ∂x j 3 k,k/= j
∂ xk 2ψ ∂ xk ∂x j
(7.30)

where ψ = (3/2)(εi /σi ); εi and σi are strain and stress intensities, respectively:
/ [
| Σ Σ
2 | ( )
m m
εi = · ] eii − e j j + 6 ei2j ;
3 j,i i/= j j,i i/= j
[
| Σ Σ
1 | ( )2
m m
σi = √ ] σii − σ j j + 6 σi2j .
6 j,i i/= j j,i i/= j

The boundary conditions of the additional loading method have the following
form:

Σ
m
∂u j (y) Σ
m
∂u k
λθ(y)n j + μ nk + μ n k = X ν j + X ν(0)j , (7.31)
k=1
∂ xk k=1
∂x j

where
172 7 Mathematical Models of Plasticity Theory

m (
Σ )[( ) ] ( )[ ]
1 ∂u j ∂u k 1 ∂u j 1
X ν(0)j = μ− + n k + 2μ − − θ n j.
k,k/= j
2ψ ∂ xk ∂x j ψ ∂x j 3
(7.32)

According to (7.29) and (7.31), the initial elastic–plastic problems can be reduced
to a sequence of “pseudo-problems” for solving the theory of elasticity.
The procedure for solving the “pseudo-problems” of theory of elasticity is the
implementation of the methods of successive approximations, taking in the first
approximation that all additional volumetric and surface loads are zeros (that
is, ψ = (1/2)μ). After solving the theory of elastic problem for given forces
X (k) (k)
j , X ν j ( j = 1, . . . , m), we find displacements in the kth approximation u j
(k)

. Then we determine strains e(k) (k)


j and strain intensity ε j . And from the given defor-
( )
(k)
mation diagram, taking into account hardening, we determine σik = . εi and
( )
(k) (k)
then ψ(k+1) = (3/2)εi /. εi . By formulas (7.30) and (7.32), we find additional
loads X (0) (0)
j , X ν j and again solve the theory of elastic problem. The solution must
be continued until two consecutive approximations differ among themselves by an
infinitesimal predetermined value.
Let’s consider an approach based on the method of additional deformations.
The transformed continuity equations in the case of solving elastic–plastic
problems by the method of additional deformations have the form:
[( )]
ν Σ ∂ X k
m
1 ∂ 2 σ0 ∂ Xi ∂Xj
.σi j + =− + − δi j + X i(0)
j ,
1 + ν ∂ xi ∂ x j ∂x j ∂ xi 1 − ν k=1 ∂ xk
(7.33)

where
⎧ [( ) ]
∂2 1 ( σ0 )
X ii(0)= 2μ ψ− σkk −
∂ x 2j 2μ 3
[( ) ] [( ) ]⎫
∂2 2 ( σ0 ) ∂2 1
+ ψ − σ jj − − 2ψ − σk j ; (7.34)
∂ xk2 μ 3 ∂ xk ∂ x j μ
⎧ 2 [( ) ] [( ) ]
∂ 1 ∂2 1
X i(0) = μ 2ψ − σ i j + 2ψ − σk j
j
∂ xk2 μ ∂ xk ∂ xi μ
[( ) ] [( ) ]⎫
∂2 1 ∂2 1 ( σ0 )
+ 2ψ − σki − 2 ψ− σkk − ,
∂ xk ∂ x j μ ∂ xi ∂ x j 2μ 3
(i, j, k = 1, . . . , m; ) i /= j /= k. (7.35)
7.10 Plastic Flow Theory 173

As in the previous case, the initial problem is given to the sequence of the theory
of elastic problems based on the method of successive approximations, in the first
approximation believing ψ = (1/2)μ.
The components of the strain tensor in the k th approximation εi(k)
j are determined
(k)
through the k approximate solution σi j by the following formulas:
th

[ ]
Σ ( )
εii(k) = (1/E) σii(k) −ν σll + (ψ − (1/2μ)) σii(k) − σ0 ;
l,l/=i
( )
εi(k)
j = (1/μ)σi(k) j + (2ψ − (1/μ))σi(k)
j .

Then we calculate the function ψ, and by the formulas (7.34) and (7.35) we
find X ii(0) and X i(0)
j . Then we solve the theory of elastic problems with new body
forces. The calculations continue until the difference between the two consecutive
approximations is sufficiently small.

7.10 Plastic Flow Theory

For calculations in the case of large plastic deformations, the theory of plastic flow
is used. Its main difference from deformation theory is that it is accepted that there is
no unambiguous connection between stresses and plastic strains under both simple
and complex loadings.
Stresses at plastic deformations in final state depend on deformation path. In this
regard, the equations describing plastic deformation cannot in principle be finite
relations linking the components of stresses and strains, but should be differential
relationships.
The equations of the plastic flow theory establish a connection between infinites-
imal increments of strains and stresses, the stresses themselves and some parameters
of the plastic state.
Remark The physical equations corresponding to this theory for the flat problem
were first obtained by Saint-Venan, and for the spatial problem—M.K. Levi5 and
later Mises.
In flow theory, the plastic deformation of material is likened to a viscous fluid
flow.
As noted earlier, the transition to the plastic state in the vicinity of the body point
is determined by the equation f (σij ) = 0, which in the six-dimensional stress space
describes the surface of yield (loading). If the material is hardened, the yield surface
changes with increasing plastic strains, its equation contains the hardening parameter
η.
The following hypotheses underlie the theory of plastic flow.

5 Maurice Lévy (1838–1910), French mathematician, mechanic and engineer.


174 7 Mathematical Models of Plasticity Theory

1. Volumetric strains are elastic. As in the theory of small elastoplastic deforma-


tions (7.6), the ratio is fulfilled:

σ = 3K ε.

That is, the volume relative change is an elastic strain proportional to the average
stress:

θ = σ0 /K ; d θ = 3(d ε0e + d ε0 p ) = d σ0 /K . (7.36)

In this case, the proportional coefficient K is the same as in the ratios used when
the behavior of the material is within the elastic range:

K = E/3(1 − 2ν),

3 Σ Σ
where σ0 = 1
3 Σ σii is the average stress; ε0e = 1
3
εiie , ε0 p = 1
3
εii p ; εiie and
i=1 i i
εii p are, respectively, elastic and plastic parts of general strains εii .
We remind that during plastic deformations, the body volume practically does not
change. Therefore, the plastic strain increment tensor is a deviator. Then:

d εop = 0, d εoe = d σ0 /3K .

That is, the material is incompressible in the plastic state:


p p p
ε p = (ε11 + ε22 + ε33 )/3 = 0.

Remark This incompressible condition can also be written as equal to zero the rate
of volumetric plastic strain, i.e., d ε p /d t = 0.
2. Gradient hypothesis. Strain increment vector is directed perpendicular to
yield surface. This is equivalent to the assumption of the proportionality of
strain component increment and the gradient vector component to the yield
surface (partial derivatives of the surface equation over the corresponding stress
components):

p ∂f p ∂f
d ε11 = d λ , . . . , 2d ε13 = d λ . (7.37)
∂ σ11 ∂ σ13

Ratios (7.37) express the law of flow associated with the accepted plasticity
condition f = 0,where d λ is a differential-small multiplier.
By definition, we have:
7.10 Plastic Flow Theory 175

Fig. 7.16 Yield surface

p
dU = σi j d εi j .

Using the ratio (7.37), we obtain:

∂f
dU = d λσi j = d λ σ grad f, (7.38)
∂ σi j

where σ is the radius-vector of the point in the six-dimensional stress space corre-
sponding to the components σij . The part of the yield surface and the vectors consti-
tuting the scalar product in the formula (7.38) is showed in Fig. 7.16. Therefore, the
multiplier d λ is proportional to the stress working density on plastic strains.
The function f is called the plastic potential, since the increment of plastic strains
is determined through the derivative function f by the corresponding arguments
(7.38). The yield surface, or plastic potential, is determined experimentally. One of the
important requirements for the yield surface construction is the hardening criterion
formulated by D. Drucker. The hardening criterion formulated by D. Drucker is as
follows.
Let there be a certain stress state σ* to which the yield surface S * corresponds
(Fig. 7.17). The yield surface divides the six-dimensional stress space into two areas.
Transition from a point σ* to a region of its positive values is accompanied by active
plastic deformation and an increment of the vector of plastic deformation. Transition
from the end of the vector σ* to the area of negative values corresponds to unloading
subject to elastic laws; plastic deformation is unchanged.

Fig. 7.17 Yield surfaces


during loading
176 7 Mathematical Models of Plasticity Theory

Let’s now consider the state defined by vector σ, the end of which lies in a positive
region. This state corresponds to the yield surface S.
The Drucker postulate claims the non-negativity of the stress increment work on
real deformation displacements of per cycle, when the state from the point σ* along
some path passes to σ, then returns to σ* :

(σi j − σi∗j )d εi j ≥ 0. (7.39)

In the theory of plastic flow, it is assumed that for a given material the stress
intensity is a function of the integral from the intensity of the plastic strain increments:

σi = F( d εi p ). (7.40)

Function F is determined by the diagram of material tension. To do this, you must


first convert the function σ = f (ε) to a function σ = f ∗ (ε p ), (Fig. 7.18).
Indeed, for uniaxial tension, we have:

σx = σ y = 0; σz = σ; τx y = τ yz = τzx = 0;

d εx p = d ε yp = − 2 p ; d εx p + d ε yp + d εzp = 0; d εzp = d ε p .

Then, with due regard for:

1 Σ
σi = √ (σ1 − σ2 )2 + (σ2 − σ3 )2 + (σ3 − σ1 )2 ,
2

D1 D
C1 C
B1 B

А1 А

Fig. 7.18 Material tensile diagram


7.10 Plastic Flow Theory 177
√ /
2 3 [( )2 ( )2 ( )2 ]
d εip = (d εxp − d εyp )2 + (d εyp − d εzp )2 + (d εzp − d εxp )2 + d γxyp + d γyzp + d γzxp ,
3 2

we get:

σi = σ , d εi p = d ε p . (7.41)

The curve A1 B1 C 1 D1 expresses the relationship between stress intensity and the
Odquist parameter, i.e., σi = F(q).

Remark Odqvist parameter q characterizes accumulated plastic strain: q = d εi p .

The increment of the plastic strain components can be expressed as follows:

( ) 3 d εi p ( )
d εi j p = σi j − δi j σ0 . (7.42)
2 σi

Equation (7.42) shows that components of plastic strain increment are proportional
to components of a stress deviator.
We add to the plastic strain components defined by (7.42) the elastic strain
components defined by the formulas:
( )
( ) 1 3ν
εi j e = σi j − σ0 δi j .
2μ 1+ν

As a result, we obtain formulas for determining the components of complete strain


increments:
[ ]
1 3ν 3 d εi p ( )
d εi j = d σi j − δi j d σ0 + σi j − δi j σ0 . (7.43)
2μ 1+ν 2 σi

Equation (7.43) is the basic equation of plastic flow theory and it is called
Prandtl-Reis equation. The relationship between stress intensity and strain incre-
ment intensity is taken as (7.40).
If in the Prandtl-Reis equations, we neglect the components of elastic strains
(which are permissible for developed plastic deformation), then we get the equation
of the Saint-Venan-Mises plasticity theory:

3 d εi ( )
d εi j = σi j − δi j σ0 . (7.44)
2 σi

Dividing both parts of the Eq. (7.44) by dt, and taking into account d εi /dt = ξi ,
we obtain the physical equation of the relationship between the strain rate and the
components of the stress deviator:
178 7 Mathematical Models of Plasticity Theory

3 ξi ( )
ξi j = σi j − δi j σ0 , (7.45)
2 σi

where ξi is the intensity of strain rates determined by the formula:


√ /
2 ( )2 ( )2 3( )
ξi = ξx − ξ y + ξ y − ξz + (ξz − ξx )2 + η2x y + η2yz + η2zx .
3 2

Obviously, strain rates ξi j are not unambiguously determined when setting the
stresses. When setting the strain rates ξi j , the components of stress deviator si j are
determined uniquely. It is easy to verify that the components si j defined by formulas
(7.45) (because si j = σi j − δi j σ0 ) are identical to the Mises yield condition. It should
also be noted that in the yield state, the uncertainty of the strain rate component is
necessary to be able to meet the strain compatibility conditions.
The Prandtl-Reis Eq. (7.43) associates stress components with infinitesimal incre-
ments of stress and strain components, i.e., they are not finite relations (unlike the
deformation theory of plasticity).
The ratios (7.43), generally speaking, are not integrated, i.e., in other words, are
not reduced to finite ratios between stress and strain components. This mathematical
fact reflects the O1 (with stresses σi j ) of two paths I and II, then the components of
the strains at the point O1 according to the equations of the plastic flow theory will
be different (Fig. 7.19).
Equation (7.43) does not contain time. However, by dividing them into dt, you can
formally switch from increments d εi j to strain rates ξi j . Then the resulting equations
will look like viscous fluid flow equations. This analogy to some extent justifies
the name of the theory of plastic flow. It should be emphasized that the variable t
can be understood as a time or a monotonically increasing loading parameter, or a
monotonically increasing value (For example, the characteristic size of the plastic
zone).
At the same time, the obtained equations of the theory of plastic flow are funda-
mentally different from the equations of viscous flow. In equations of plastic flow,
unlike the latter, you can always discard dt and return to formula (7.43) that does not
contain time. In case of hardening, it is possible to calculate strains when setting the

Fig. 7.19 Various I


deformation ways upon
transition from a point O to
point O1
О1

О
I
7.11 Relationship Between Plastic Flow and Theory of Plasticity 179

loading path (i.e., when specifying σi j = σi j (t), where t is a parameter (for example,
time). You can also find stresses if you specify a strain path, i.e., εi j = εi j (t).
The equations of the Saint-Venan-Mises plasticity theory have a much simpler
structure and are finite dependencies between stress components and strain rates.
The complete system of equations in solving the problems of plastic flow
theory includes: equilibrium equations, Cauchy ratios, strain compatibility
equations and boundary conditions.

σi j , j +ρ Fi = 0; εi j = (u i , j +u j ,i )/2;
εαα ,ββ + εββ ,αα − 2 εαβ ,αβ = 0, (εαβ ,γ − εβγ ,α + εγα ,β ),α − εαα ,βγ = 0;
u i = u i0 (x) on Su , σi j l j = Ri on Sσ .

In the theory of plastic flow, the theorem on the uniqueness of fields of increments
of stresses, strains, and displacements in a strengthening body has been proved. It
is impossible to guarantee the singular increments of strains and displacements in
the case of non-hardening materials. Therefore, obviously, the equations of flow
theory are more complex than the equations of the theory of small elastoplastic
deformations. However, it has been shown that with simple loading, both theories
give the same solution. In the case of complex loading, the results, obtained using
the theory of plastic flow, are better aligned with experimental data. In more details,
the theory and problems of the plastic flow are considered in numerous monographs
and articles.

7.11 Relationship Between Plastic Flow and Theory


of Plasticity

Theory of plastic flow and deformation theory of plasticity coincide only in the case
of simple loading. In the case of complex loading, these theories lead to different
results.
Note an important case for applications when the results obtained from two
theories are close to each other.
As it is known, in deformation spaces, the deformation path is depicted as a
line (Fig. 7.20).
Let, starting from a point, the deformation path approach to a straight line (dashed
line). In this case, we will say that deformation develops in a certain direction. If
this case occurs, then the stress states calculated according to both theories come
closer. At the same time, the influence of the history of complex deformation rapidly

Fig. 7.20 Deformation path


180 7 Mathematical Models of Plasticity Theory

weakens and a constant stress state is established, which determined by those fixed
deformation rates that are specific to the rectilinear section.
Experiments confirm the theory of plastic flow much more fully than the defor-
mation theory. Experimental data also indicate the coincidence of the directions of
the principal axes of the stress tensor and the plastic strain increment tensor.
It should be noted, however, that under complex loading, especially with inter-
mediate unloading, there is a marked influence of the anisotropy that the material
acquires during the plastic deformation.

7.12 Formula and General Methods

7.12.1 Formula of the Problems of Plasticity Theory

Summarizing the material presented in the previous sections concerning the setting
of problems of plasticity theory.
The problems of elastoplasticity usually include cases where the elastic and plastic
deformations that occur in the body during loading have almost the same order. The
setting of such problems should take into account ratios in elastic and plastic zones
and at the boundary between them.
We consider a solid deformable body under the influence of surface (X ν , Yν , Z ν )
(including reactions) and volumetric (X, Y, Z ) forces. Elastoplastic properties of
materials are given by deformation diagram (σi − εi ). The task is( to) define 17
unknown coordinate( functions
) (x, y, z): six stress tensor components σi j ; six strain
tensor components εi j ; three components of the deformation vector (u, v, w); stress
intensity (σi ) (one); strain intensity (εi ) (one). To determine these functions in the
framework of plasticity theory, there is the following system of equations.
1. Equations arising from the consideration of the static equilibrium state of the
solid:
(a) Differential equilibrium equations:

Σ
3
∂σi j
+ X i = 0, (7.46)
j=1
∂x j

(b) Boundary conditions:

Σ
3
σi j n j = X νi . (7.47)
j=1
7.12 Formula and General Methods 181

2. Equations arising from considering the geometry of a problem:


(a) Dependencies between strain components and deformation vector components
(Cauchy equations)


∂v ⎪
εx = ∂∂ux , γx y = ∂u ∂y
+ ∂x ⎬
..................................... , (7.48)

..................................... ⎭

(b) Saint-Venan compatibility equations (as a result of Eqs. (7.48))

∂2ε ∂2γ

∂ 2 εx ⎪
∂ y2
+ ∂ x 2y = ∂ x∂xyy ⎪



............................. ⎪



( ............................
) ( 2 )
∂γ yz ∂γzx ∂γx y . (7.49)

− + + = 2 ∂ εx ⎪

∂x ∂x ∂y ∂z ∂ y∂ z ⎪


......................................................... ⎪



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

Remark As can be seen from the above equations, the static Eqs. (7.46) and (7.47)
and the geometric relations (7.48) and (7.49) in the theory of plasticity have the
same form as it in the theory of elasticity. The other equations are only the consti-
tutive equations, i.e., the relations between stresses and strains. Therefore, when
solving elastoplastic problems, it is necessary to write physical constitutive equations
according to one of the plasticity theories instead of the Hooke’s law.
3. Equations arising from consideration of the mechanical behavior of materials,
i.e., constitutive equations between stresses and strains:
(a) On the theory of small elastoplastic deformations

σi

σx − σ0 = 2 3ε i(
(εx − ε0 ), ⎪

2 σi
) ⎪

σ y − σ0 = 3 εi ε y − ε0 ,
(7.50)
σz − σ0 = 23 σεii (εz − ε0 ), ⎪



τx y = 3εσii γx y , τ yz = 3εσii γ yz , τzx = σi
3εi
γzx .

or

εx − ε0 = 2σ 3εi
i(
(σx − σ0 ), ⎪

) ⎪

ε y − ε0 = 2σi σ y − σ0 ,
3εi
, (7.51)
εz − ε0 = 2σ3εi
(σz − σ0 ), ⎪

i ⎪

γx y = σi τx y , γ yz = 3εσii τ yz , γzx =
3εi 3εi
σi
τzx
182 7 Mathematical Models of Plasticity Theory

where σi = .(εi ), σi , εi are stress and strain intensities.


(b) On the theory of plastic flow

⎧ [ ( )] 3 d εi p

⎪ d εx = E [d σx − μ d σ y + d σz ] + 2 σi ((σx − σ0))
1



⎪ d ε y = E1 d σ y − μ(d σx + d σz ) + 23 σii p σ y − σ0


⎪ [ ( )]
⎨ dε
d εz = E1 d σz − μ d σx + d σ y + 23 σii p (σz − σ0 )
, (7.52)


dτ dε
d γx y = Gx y + 3 σii p τx y



⎪ .......................................................................



.......................................................................
(∫ )
where σi = . d εi p ,
[( )2 ( )2 ( )2
√ | |
2 | d ε⎧xp[ − d εyp ] + [d εyp − d]εzp + d εzp −⎫ d εxp
d ε ip = ] 2 ( )2 2 ( ) 2 [ ]2 .
3 + 3 d γxy p + d γyz p + d(γzx ) p

The use of physical equations according to the theory of plastic flow in the form
(7.52) in solving elastoplastic problems is associated with great mathematical diffi-
culties, since they are non-linear and have a rather complex structure. Therefore, when
solving problems in which significant plastic deformations are taken into account in
comparison with elastic ones, the components of elastic strains are neglected and use
the Saint-Venan-Mises equations, which for a rigid plastic body have the form:
⎫ ⎧
3 ξi
ξx = 2 σi (σ
− σ0 ) ⎪ ⎪
⎨ ηx y =
3ξi
τ
3 ξi
( x )⎬ σi x y
ξy = 2 σi
σ y − σ , η yz = 3ξi
τ
σi yz . (7.53)
3 ξi
0

⎭ ⎪
⎩η =
ξz = 2 σi (σz
− σ0 ) zx
3ξi
τ
σi zx

/( )2 ( )2

ξx (− ξ y + ξ y − ξ)z + (ξz − ξx )2
where σi = .(ξi ); ξi = 2
.
3
+ 23 ηx2 y + η2yz + ηzx
2

When solving elastoplastic problems using physical Eq. (7.53) according to the
theory of plastic flow, it is necessary to write Cauchy Eq. (7.48) and, therefore, strain
compatibility equations respectively in the following form:

∂ 2 ξx ∂ 2ξy ∂ 2 ηx y
+ = ,
∂ y2 ∂x2 ∂ x∂ y
∂ 2ξy ∂ 2 ξz ∂ 2 η yz
+ = ,
∂ z2 ∂ y2 ∂ y∂z
∂ 2 ξz ∂ 2 ξx ∂ 2 ηzx
+ = ,
∂x 2 ∂z 2 ∂ x∂z
7.12 Formula and General Methods 183
( )
∂ ∂η yz ∂ηzx ∂ηx y ∂ 2 ξx
− + + =2 ,
∂x ∂x ∂y ∂z ∂ y∂ z
( )
∂ ∂η yz ∂ηzx ∂ηx y ∂ 2ξy
− + =2 ,
∂y ∂x ∂y ∂z ∂z∂ x
( )
∂ ∂η yz ∂ηzx ∂ηx y ∂ 2 ξz
+ − =2 .
∂z ∂x ∂y ∂z ∂ x∂ y

where
⎧ ∂v
ξx = ∂v∂x
x
, ξ y = ∂ yy , ξz = ∂v
∂z
z
,
∂vx ∂v y ∂vz ∂v y ∂vx ∂vz
ηx y = ∂y
+ ∂ x , η yz = ∂ y + ∂ z , ηzx = ∂z
+ ∂x
.

These systems of equations are complete systems of equations for solving elasto-
plastic problems with active deformation and simple loading, or loading close to
simple loading.

7.12.2 General Methods of Solving the Problems of Plasticity


Theory

As in the theory of elasticity, the problems of plasticity theory can be solved in


displacements or stresses, as well as in a mixed way. For most practical problems,
solving in a closed form is difficult due to the non-linearity of the governing partial
differential equations. Therefore, to solve non-linear equations of plasticity theory,
in most cases, various versions of the method of successive approximations are used.
Solving the problems of plasticity theory is usually reduced to solve a sequence of
linear problems, each of which can be interpreted as a certain problem of elasticity
theory (a method of elasticity solutions). Such an idea was first applied by A.A.
Ilyushin, and then developed by I.A. Birger6 and others.
Summary of the most common methods for solving problems of plasticity theories
are as follows.
Method of Additional Loads
We take the Genki-Nadai equations as basic.

εx − ε0 = .(σ ( x − σ0 ),
) ⎪


ε y − ε0 = . σ y − σ0 ,
, (7.54)
εz − ε0 = .(σz − σ0 ), ⎪


γx y = 2.τx y , γ yz = 2.τ yz , γ yz = 2.τ yz ,

where in the elastic area: . = 1/2G; in the elastoplastic region: . = 1/2G + ϕ.

6Birger Isaac Aronovich (1918–1993), Soviet mechanical scientist in the field of strength and
dynamics of aircraft engines for aviation and space purposes, doctor of technical sciences, professor.
184 7 Mathematical Models of Plasticity Theory

Let’s accept that 1/. = 2G −(2G − 1/.). In addition, we will take into account
ε0 = θ/3 , σ0 = 3K ε0 , K = E/(3(1 − 2μ)), λ = μ E/(1 − 2μ)(1 − μ). Then
the Eq. (7.54) are rewritten in the following form:
( ) ⎫
σx = 2G εx + λθ − (2G − .1 )((εx − ε0 ), ⎪
) ⎪

σ y = 2G ε y + λθ − (2G − . ) ε y − ε0 , ⎪
1 ⎪



σz = 2G εz + λθ − (2G − . )(εz − ε0 ),
1
. (7.55)
τx y = G γx y − ( G − 2.
1
γx y , ⎪

) ⎪

τ yz = G γ yz − (G − 2. ⎪
)γ yz ,
1



τzx = G τzx − G − 2. γzx .
1

Equations (7.55) differ from the Hooke’s law by additional terms. With . =
1/2G, these equations express the generalized Hooke’s law and determine the stresses
in the elastic body:
⎫ ⎫
σ∗x = 2G εx + λθ, ⎬ τ∗x y = G γx y , ⎪

σ∗y = 2G ε y + λθ, τ∗yz = G γ yz , . (7.56)
⎭ ⎪
σ∗z = 2G εz + λθ, τ∗zx = G γzx . ⎭

Enter such expressions for stresses:



σx = σ∗x − σ(0) x , ⎪⎪

................... , ⎪⎪



................... ,
∗ (0) ⎪, (7.57)
τx y = τx y − τx y , ⎪

................... , ⎪⎪



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

where σ(0) (0)


x , . . . , τx y , . . . are additional stresses:

( ) ⎫
σ(0)
x = 2G − . (εx − ε0 ), ⎪
1


.................................... , ⎪




...................................
( ) ,
(0) , (7.58)
τx y = G − 2. γx y , ⎪
1


................................. , ⎪⎪



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

There is such a relationship between stress intensities in the elastic body and
additional stresses, as it shown in Fig. 7.21:

σi = σi∗ − σi(0) . (7.59)


7.12 Formula and General Methods 185

Fig. 7.21 To the additional


loads method B

The physical meaning of equality (7.59) is clear from the generalized diagram
(see Fig. 7.21). According to the Fig. 7.21:
( )
1
σi(0) = σi∗ − σi = εi (tg β − tg α) = 3 εi G− . (7.60)
2.

Assuming that the material is elastic, using a generalized deformation diagram at


a given strain intensity εi , we find a point B.
The correction σi(0) “returns” the calculated point to the deformation curve (point
A). The stresses σx , σ y , . . . ., τzx must satisfy the Eq. (7.46). These equations, taking
into account (7.57), have the form:
∂τ∗

∂σ∗x ∂τ∗
∂x
+ ∂ yx y + ∂ zx z + X + X 0 = 0, ⎪


∂σ∗y ∂τ∗x y ∂τ∗yz
+ + + Y + Y 0
= 0, , (7.61)
∂y ∂x ∂z ⎪

∂τ∗x z
+
∂τ∗yz
+
∂σ∗z
+ Z + Z = 0,
0 ⎭
∂x ∂y ∂z

where X, Y, Z are components of the body force; X 0 , Y 0 , Z 0 are components of


additional body force:
⎧ ( (0) (0) (0)
)


⎪ 0 = − ∂σx + ∂τx y + ∂τx z =

⎪ X

⎪ ∂x ∂y ∂z

⎪ ⎧ [( ) ] [( ) ] [( ) ]⎫

⎪ ∂ 1 ∂ 1 ∂ 1

⎪ − 2G − (ε − ε ) + G − γ + G − γzx

⎪ ∂x ( .
x 0
∂y 2.
xy
∂z 2.

⎪ )

⎪ (0)
∂τ yx
(0)
∂σ y ∂τ yz
(0)


⎨ Y0 = − + + =
∂x ∂y ∂z (7.62)
⎪ ⎧ [( ) ] [( ) ] [( ) ]⎫

⎪ ∂ 1 ∂ 1 ( ) ∂ 1

⎪ − ∂ x G − 2. γ yx + ∂ y 2G − . ε y − ε0 + ∂z G − 2. γ yz



⎪ ( (0) (0) (0)
)



⎪ 0 = − ∂τzx + ∂τzy + ∂σz =

⎪ Z

⎪ ∂x ∂y ∂z

⎪ ⎧ [( ) ] [( ) ] [( ) ]⎫

⎪ ∂ 1 ∂ 1 ∂ 1
⎩− G− γzx + G− γzy + 2G − (εz − ε0 ) .
∂x 2. ∂y 2. ∂z .
186 7 Mathematical Models of Plasticity Theory

Stresses must satisfy the boundary conditions (7.47) on the body surface.
Expression (7.47), taking into account (7.57), is written in the form:

σ∗x l + τ∗x y m + σ∗x z n = X v + X ν(0) ⎪

τ∗yx l + σ∗y m + τ∗yz n = Yv + Yν(0) , (7.63)

τ∗zx l + τ∗zy m + σ∗z n = Z v + Z ν(0) ⎭

where X v , Yv , Z v are surface load components; X v(0) , Yv(0) , Z v(0) are components
of additional surface loads:
⎧ [( ) ] ( )
⎪ (0) (0) (0) (0)
X v = σx l + τx y m + τx z n = 2G − . 1 (ε − ε ) l + G − 1 (γ m + γ n ),

⎪ x 0 xy xz
⎨ ( ) [( 2. ) ]
(0) (0) (0) (0)
Yv = τ yx l + σ y m + τ yz n = G − 2.1 (γ l + γ n ) + 2G − 1 (ε − ε ) m, (7.64)
yx yz . ) y 0

⎪ ( ) [( ]

⎩ Z v(0) = τ(0) (0) (0) 1 ( ) 1
zx l + τzy m + σz n = G − 2. γzx l + γzy m + 2G − . (εz − ε0 ) n.

Using Eqs. (7.56), (7.57), Cauchy Eqs. (7.48), after some transformations, the
differential equilibrium Eqs. (7.61), taking into account (7.62), are written in the
following form:
⎧ {( )[ ( )]}

⎪ (λ + G) ∂∂θx + G.u + X = ∂∂x 2G − .1 ∂∂ux − 13 ∂∂ux + ∂v + ∂w +

⎪ [( ( )] ∂ y ∂ z

⎪ ) [( )( )]

⎪ + ∂∂y G − 2. 1 ∂u
+ ∂∂vx + ∂∂z G − 2. 1 ∂u
+ ∂w

⎪ ∂y {( [ ( ∂z ∂x )]}

⎪ )
⎨ (λ + G) ∂θ + G.v + Y = ∂ 2G − 1 ∂v − 1 ∂u + ∂v + ∂w +
∂y [ ( ∂ y )] . [( ∂ y 3 ∂
( x ∂ y )]∂ z
( ) )

⎪ + ∂∂x G − 2. 1 ∂u
+ ∂∂vx + ∂z ∂
G − 2. 1 ∂v
+ ∂w

⎪ ∂y {( [ ( ∂z ∂y )]}

⎪ )

⎪ (λ + G) ∂θ
+ G.w + Z = ∂
2G − 1 ∂w
− 1 ∂u
+ ∂v
+ ∂w
+

⎪ ∂z ∂z .[ ∂z 3 (∂ x ∂ y )] ∂z

⎪ [( )( )] ( )
⎩ + ∂∂x G − 2. 1 ∂u
+ ∂w + ∂∂y G − 2. 1 ∂w
+ ∂v .
∂z ∂x ∂y ∂z
(7.65)

The system of Eqs. (7.65) is a synthesis of static, geometric, and physical equations
of the original problem. The Eqs. (7.65) differ from the Lamé equations in the theory
of elasticity in the presence of additional terms on the right side.
Similarly, you can convert boundary conditions (7.63), taking into account (7.64),
have the form:
( ) ( )
∂u ∂u ∂u ∂u ∂v ∂w
λθl + G l+ m+ n +G l+ m+ n = Xν
∂x ∂y ∂z ∂x ∂x ∂x
( )[( ) ( ) ] ( )[ ( )]
1 ∂u ∂v ∂u ∂w 1 ∂u 1 ∂u ∂v ∂w
+ G− + m+ + n + 2G − − + + l,
2. ∂y ∂x ∂z ∂x . ∂x 3 ∂x ∂y ∂z
( ) ( )
∂u ∂u ∂u ∂u ∂v ∂w
λθ l + G l+ m+ n +G l+ m+ n = Xν
∂x ∂y ∂z ∂x ∂x ∂x
( )[( ) ( ) ]
1 ∂u ∂v ∂u ∂w
+ G− + m+ + n
2. ∂y ∂x ∂z ∂x
7.12 Formula and General Methods 187
( )[ ( )]
1 ∂u 1 ∂u ∂v ∂w
+ 2G − − + + l,
. ∂x 3 ∂x ∂y ∂z
( ) ( )
∂v ∂v ∂v ∂u ∂v ∂w
λθ m + G l+ m+ n +G l+ m+ n = Yν
∂x ∂y ∂z ∂y ∂y ∂y
( )[( ) ( ) ]
1 ∂u ∂v ∂v ∂w
+ G− + l+ + n
2. ∂y ∂x ∂z ∂y
( )[ ( )]
1 ∂v 1 ∂u ∂v ∂w
+ 2G − − + + m,
. ∂y 3 ∂x ∂y ∂z
( ) ( )
∂w ∂w ∂w ∂u ∂v ∂w
λθ n + G l+ m+ n +G l+ m+ n = Zν
∂x ∂y ∂z ∂z ∂z ∂z
( )[( ) ( ) ]
1 ∂u ∂w ∂w ∂v
+ G− + l+ + m
2. ∂z ∂x ∂y ∂z
( )[ ( )]
1 ∂ w 1 ∂u ∂v ∂w
+ 2G − − + + n. (7.66)
. ∂z 3 ∂x ∂y ∂z

Equations(7.65) together with Eq. (7.66) allow to solve the problem of plasticity
theory in displacements.
If we assume that the terms arising from the presence of additional terms in
(7.55) and transferred to the right parts (7.65) and (7.66) are known, then we get a
system of equations of(the theory of elasticity ) with respect
( to displacements,
) but with
additional volumetric X (0) , Y (0) , Z (0) and surface X v(0) , Yv(0) , Z v(0) forces. In the
first approximation, we assume that all additional volumetric and surface loads are
zeros (i.e., . = 1/2G). Then we have the “usual” problem of the theory of elasticity
in displacements. Equations (7.65) are converted to the Lamé equations of the theory
of elasticity, and Eqs. (7.66) to the boundary conditions of the theory of elasticity in
displacements.
Let the specified problem of the theory of elasticity for given forces
( Y, Z ; X v , Yv , Z v be solved
X, ) and we obtained displacements (u 0 , v0 , w0 ), strains
εx0 , ε y0 , εz0 , γx y0 , γ yz0 , γzx0 , and intensity of strains (εi0 ). We determine σi0 =
εi0
.(εi0 ), and then .1 = 23 .(ε i0 )
from the given deformation diagram (Fig. 7.23),
taking into account the hardening. We know the displacements and value .1 , and
now we obtain additional loads X (0) , Y (0) , Z (0) ; X v(0) , Yv(0) , Z v(0) (which in this case
are different from zero), using (7.62) and (7.64). Again, we solve the problem of
the theory of elasticity with additional loads and define u 1 , v1 , w1 , and, therefore,
εx1 , ε y1 , εz1 , γx y1 , γ yz1 , γzx1 and εi1 . From the deformation diagram at the calculated
value εi1 , taking into account hardening, we obtain σi1 = .(εi1 ). Then we determine
εi1
.2 = 23 .(ε i1 )
again, etc.
The solution must continue until the current approximation differs from the
previous one by an infinitesimal predetermined value.
Let’s consider the implementation of the method of elastic solutions in the form
of the method of additional loads in a slightly different form
We use the relationship between stress and strain intensities in the form proposed
by A.A. Ilyushin. (i.e. σi = 3G εi [1 − ωi (εi )]). Then the Eq. (7.55) take the form:
188 7 Mathematical Models of Plasticity Theory
( )
σx = 2G ∂∂ux + 2G ωi( 3θ − ∂u
∂x )
,
σ y = 2G ∂v
∂y
+ 2G ωi θ
3
− ∂v
∂y
,
. . . . . . . . . . . . . (. . . . . . . . ) (7.67)
τx y = G(1 − ωi ) ∂u ∂y
+ ∂∂vx ,
.....................

Now equilibrium Eq. (7.46), taking into account (7.67) and provided that all terms
containing the function ωi are transferred to the right part, are written as:

(G + λ) ∂∂θx + G.u + X = X (0) ⎪

(G + λ) ∂∂θy + G.v + Y = Y (0) , (7.68)

(G + λ) ∂θ
∂z
+ G.w + Z = Z (0) ⎭

where
⎧ [ ( )

⎪ X (0) = G ωi .u + 13 ωi ∂∂θx + 43 ∂ω i ∂u
− 23 ∂ω i ∂v
+ ∂w +

⎪ ( ) ∂ x ∂ x ∂ x ] ∂ y ∂ z

⎪ ( )

⎪ + ∂ωi ∂u + ∂∂vx + ∂ω i ∂w
+ ∂u ;

⎪ [ ∂y ∂y ∂z ∂x ∂z

⎪ ( )
⎨ Y (0) = G ωi .v + ωi +
1 ∂θ 4 ∂ω i ∂v ∂ω
− 3 ∂ yi ∂w
2
+ ∂∂ux +
( 3 ∂ y) 3 ∂ y ( ∂ y )] ∂z
(7.69)

⎪ + ∂ωi ∂u
+ ∂v
+ ∂ωi ∂v
+ ∂w
;

⎪ [ ∂x ∂y ∂x ∂z ∂ z ∂y ( )



⎪ 1 4 ∂ω ∂w 2 ∂ω
Z (0) = G ωi .w + 3 ωi ∂ z + 3 ∂zi ∂ z − 3 ∂zi ∂∂ux + ∂v
∂θ
+

⎪ ( )] ∂ y

⎪ ( )
⎩ + ∂ω i ∂w
+ ∂u + ∂ω i ∂v
+ ∂w .
∂x ∂x ∂z ∂y ∂z ∂y

To solve Eq. (7.68), we add boundary conditions in displacements in the following


form:
⎧( ) ( ) ( )

⎪ 2G ∂∂ux + λθ l + G ∂u + ∂∂vx m + G ∂w + ∂u n = X v + X v(0) ,

⎨( ) ∂
( y ) (∂ x ∂z )

2G ∂v
∂y
+ λθ m + G ∂u ∂y
+ ∂∂vx l + G ∂v + ∂w n = Yv + Yv(0) , (7.70)

⎪ ( ) ( ) ( ∂z ∂ y)

⎩ 2G + λθ n + G
∂w ∂w ∂u ∂v ∂w
∂z ∂x
+ ∂ z l + G ∂z + ∂ y m = Z v + Z v(0) ,

where
⎧ [( ) ( ) ( ∂w ∂u ) ]

⎪ X (0)
= G ω 2 ∂u
− 2
θ l + ∂u
+ ∂v
m + + ∂z n ,

⎨ v
i
[( ∂ x 3) ( ∂y ∂x ) ( ∂x ) ]
Yv(0) = G ωi ∂u ∂
+ ∂∂vx l + 2 ∂v ∂
− 23 θ m + ∂v ∂z
+ ∂w

n , (7.71)

⎪ [( y
) ( y ) ( ∂w 2 ) ]
y
⎪ (0)
⎩ ∂w ∂u ∂v ∂w
Z v = G ωi ∂ x + ∂ z l + ∂ z + ∂ y m + 2 ∂z − 3 θ n .

Thus, the solution of the problem of the theory of plasticity under active defor-
mation and simple loading, taking into account the Ilyushin dependence σi =
3G εi (1 − ωi ), came down to solve the Eqs. (7.68) under boundary conditions (7.70).
7.12 Formula and General Methods 189

This method of elastic solutions of A.A. Ilyushin is also based on the principle
of successive approximations (allocation procedure). In the first approximation, it
is accepted that ωi0 = 0. In this case, according ( to (7.69) and (7.71), all additional )
volumetric and surface forces are zeros X v(0) = Yv(0) = Z v(0) = X (0) = Y (0) = Z (0) .
Equations (7.68) are Lamé equations of the theory of elasticity. Therefore,
the solution in the first approximation is to solve the problem of the theory
of elasticity. Supposing that the solution to this problem at given forces
X, Y, Z ; X v , Yv , Z v is u, v, w. Using (7.67), we obtain stresses as a function of
coordinates σx , σ y , σz , τx y , τ yz , τzx , and, hence, stress intensity σi . From the defor-
mation diagram, we determine εi , and from the analytical dependence of A.A.
Ilyushin we obtain ωi as a function of coordinates. Now by formulas (7.69) and
(7.71) we find additional loads X (0) , Y (0) , Z (0) ; X v(0) , Yv(0) , Z v(0) as coordinate func-
(tions. After that, we solve the) problem of the theory of elasticity for volumetric
X − X (0) , Y − Y (0) , Z − Z (0) and surface X v + X v(0) , Yv + Yv(0) , Z v + Z v(0) forces.
The solution to the problem of the theory of elasticity in this case is displacements
u 2 , v2 , w2 in the second approximation, etc. The calculations can be completed when
the difference between the results of two consecutive approximations is sufficiently
small, that is, it will be within the required accuracy.
Method of additional deformations
One of the types of elastic solutions is the method of additional deformations. If in
Genki-Nadai Eqs. (7.54), we denote(. = (1/2G) )+ (. − (1/2G)), and we also take
into account ε0 = [(1 − 2μ)/(3E)] σx + σ y + σz , then after small transformations
of Eqs. (7.54), we can write these equations in the following form:
⎧ [ ( )] ( )

⎪ εx = 1
σ
E[ x
− μ σ y + σz + . − 2G 1
(σx − σ0)),

⎪ ] ( )(

⎪ εy = 1
σ − μ(σx + σz ) + . − 2G 1
σ − σ0 ,

⎨ E[ y ( )] ( ) y
εz = 1
σz − μ σx + σ y + . − 2G (σz − σ0 ),
1
E ( ) (7.72)

⎪ γx y = G1 τx y + 2. − G1 τx y ,

⎪ ( )

⎪ γ yz = G1 τ yz + 2. − G1 τ yz ,

⎩ ( )
τzx = G1 τzx + 2. − G1 τzx .

These equations differ from the generalized Hooke’s law by additional terms.
If . = (1/2G) , then these equations express the generalized Hooke’s law and
determine the strains (relative deformations) in the elastic body:
⎧ [ ( )]

⎪ εx∗ = E1 [σx − μ σy + σz ] ,

⎨ ε∗ = 1 σ − μ(σ + σ ) ,
y E[ y ( x z
)] (7.73)
⎪ ∗
ε
⎪ z = 1
σ − μ σ + σ ,
⎪ E z x
⎩γ∗ = 1 τ , γ∗ = 1 τ , γ∗ =
y

xy G xy yz G yz zx
1
τ .
G zx

Then
190 7 Mathematical Models of Plasticity Theory

εx = ε∗x + ε(0) ∗ (0) ∗ (0)
x , ε y = ε y + ε y , εz = εz + εz ,
∗ ∗ ∗ (7.74)
γx y = γx y + γx y , γ yz = γ yz + γ yz , γzx = γzx + γ(0)
(0) (0)
zx .

where ε(0) (0) (0) (0) (0) (0)


x , ε y , εz , γx y , γ yz , γzx are plastic strains in elastic–plastic bodies:

⎧ (0) ( )

⎪ εx = . − 2G 1
(σx − σ0 ),

⎪ (0)
( )( )

⎪ ε = . − 1
σ y − σ0 ,
⎪ (0) (

y 2G )
εz = . − 2G 1
(σz) − σ0 ),
(0)
( (7.75)

⎪ γ = 2. − 1
τ ,


xy ( G ) xy

⎪ γ (0)
= 2. − 1
τ ,

⎩ yz ( G ) yz
(0)
γzx = 2. − G τzx .1

Strain intensity is defined as:

εi = εi∗ + εi(0) , (7.76)

where εi∗ , εi(0) are intensity of elastic and plastic strains, respectively.
The physical meaning of equality (7.76) is clear from Fig. 7.22, where:
( )
1
εi(0) = 3σi −G . (7.77)
2.

Consider the problem of plasticity theory in stresses


Differential equilibrium Eqs. (7.46) and boundary conditions (7.47) remain the same.
The strain compatibility Eqs. (7.49), due to the presence of additional (underlined)
terms in Eq. (7.72), will contain additional terms:

Fig. 7.22 To the method of


additional deformations
B А
7.12 Formula and General Methods 191

Fig. 7.23 To the difference


in the methods of solving
problems in additional loads
B А2
and additional deformations

А1

⎧ ( ) { [( )( )]
⎪ 1 ∂ 2 I 1 ( Tσ )
.σx + μ+1 = −2 ∂∂ Xx − 1−μ μ ∂ X + ∂Y + ∂ Z + 2G ∂ 2 . − 2. 1 σ y − 13 I1 ( Tσ )

⎪ ∂x2 ∂ x ∂ y ∂z ∂z 2

⎪ [( )( )] [( ) ]}


2
+ ∂ 2 . − 2. 1 2
σz − 13 I1 ( Tσ ) − ∂ ∂y∂z 2. − G1 τ yz ;

⎪ ∂y

⎪ ( ) { 2 [( )( )]

⎪ 1 ∂ 2 I 1 ( Tσ ) μ


⎪ .σ y + 1+μ = −2 ∂Y ∂X ∂Y ∂Z
∂ y − 1−μ ∂ x + ∂ y + ∂z + 2G ∂z 2 . − 2.
∂ 1 σx − 13 I1 ( Tσ )
⎪ ∂ y2 [( )( )] [( )]}

⎪ 2 2


⎪ + ∂ 2 . − 2G 1 σz − 13 I1 ( Tσ ) − ∂ ∂x∂ z 2. − G1 τzx ;

⎪ ∂x ( ) { 2 [( )( )]

⎪ 1 ∂ 2 I 1 ( Tσ ) ∂ μ ∂ ∂Y ∂ ∂

⎪ .σz + 1+μ Z X
= −2 ∂z − 1−μ ∂ x + ∂ y + ∂z + 2G Z . − 2G σx − 13 I1 ( Tσ )
1

⎪ ∂z 2 [( )( )] ∂ y 2[( ) ]}

⎪ 2 2


⎨ + ∂ 2 . − 2G 1 σ y − 13 I1 ( Tσ ) − ∂ ∂x∂ y 2. − G1 τx y ;
∂x ( ) { [( ) ] [( ) ]
.τ + 1 ∂ 2 I1 ( Tσ ) = − ∂ Z + ∂Y + G ∂ 2 − 1 τ + ∂2 − 1 τ

⎪ x y ∂ ∂z ∂z 2. yz ∂ 2. zx


1+μ y∂z [( ) ] ∂ x 2 [( 2 G )( x∂ y )]}G

⎪ + ∂2 − 1 τ − ∂ . − 1 σ − 1 I (T ) ;

⎪ ∂ x∂ z 2. G xy 2 x σ


⎪ ( ) { ∂[( y∂z
)2G ] 3 1
[( ) ]

⎪ .τ + 1 ∂ 2 I1 ( Tσ ) = − ∂ Z + ∂ X + G ∂ 2 2. − 1 τ + ∂2 2. − G1 τ yz

⎪ zx 1+μ ∂ z∂ x ∂x ∂z ∂ 2 G zx ∂ x∂ y

⎪ [( ) ] y [( )( )]}


2
+ ∂ ∂y∂ z 2. − G1 τx y − 2 ∂z∂ ∂2 . − 2G1 σ y − 13 I1 ( Tσ ) ;



⎪ ( ) { 2 [( x
) ] [( ) ]

⎪ 1 ∂ 2 I1 ( Tσ ) ∂ X ∂Y ∂ 1 ∂ 2 1

⎪ .τx y + 1+μ ∂ x∂ y = −[(∂ y + ∂ x ) + G 2. − τx y + ∂ x∂z 2. − G τ yz

⎪ ] ∂z 2 2 [( G )( )]}

⎪ ∂ 2 ∂


1
+ ∂ y∂ z 2. − G τx z − 2 ∂ x∂ y . − 2G 1 σz − 13 I1 ( Tσ ) .

(7.78)

These equations are similar to the Beltrami-Michell equations, but with additional
terms. If we put . = (1/2G) in the Eq. (7.78), then all the additional terms on the
right side will turn to zero, and we will get the usual equations of the theory of
elasticity.
Equations (7.78) together with boundary conditions (7.47) allow to solve the
problem of plasticity theory in stresses.

Remark Equations (7.78) do not integrate in closed form. To do this, various


approximate methods of solving are used.
The difference in the methods of solving problems in additional loads and addi-
tional deformations is shown in Fig. 7.23. After solving the problem of the theory of
elasticity, we “get” to the point B. Further “movement” according to the method of
additional loads is carried out in the direction to the point A1 , while according to the
method of additional deformations—in the direction to the point A2 . The convergence
192 7 Mathematical Models of Plasticity Theory

criterion of these methods is the proximity of stresses in the previous and current
approximations.

Method of elastic parameters


The physical Eqs. (7.50) and (7.51) of the theory of small elastoplastic deformations
after some transformations can be written in the form of Hooke’s law:
⎧ [ ( )]

⎪ εx = E1∗ σx − μ∗ σ y + σz ,



⎪ ....................................... ,


...................................... ,
(7.79)

⎪ γx y = G1∗ τx y ,



⎪ ...................................... ,


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

where

(σi /εi ) 1
− 1−2μ

3E (σi i )
E∗ = ; μ∗ = 2
; G ∗ = (σi /εi ). (7.80)
1+ /ε
3E (σi i )
1−2μ
1+ /ε
3E (σi i )
1−2μ

Besides,

E∗
σ0 = ε0 . (7.81)
1 − 2μ∗

As 3εi /2σi = ., therefore, formulas (7.80) can be rewritten as:

3E 1
− E (1/2.)
1−2μ
G ∗ = 1/2.; E∗ = ; μ∗ = 2
. (7.82)
2E. − 1 − 2μ 1+ 1−2μ
E
(1/2.)

Therefore, if the physical equations of small elastoplastic deformations (7.50)


and (7.51) are replaced respectively by Eqs. (7.79) and (7.81), then the solution to
the problem of plasticity theory is reduced to solving the problem of elastic theory
with parameters of elasticity theory determined by the formulas (7.80) and (7.82).
According to (7.80), the relationship between elastic parameters has the same form
for elastic constants E, G, μ , namely G ∗ = E ∗ /(2(1 + μ∗ )).

Remark This method was first proposed by prof. I.A. Birger.


The algorithm for solving the problems of plasticity theory using the method of
variable elasticity parameters is as follows.

In the first approximation we accept that . = 1/(2G) . Therefore, the elastic


parameters equal to the elastic constants: E ∗ = E, G ∗ = G, μ∗ = μ. The
physical equations in this case turn to the generalized Hooke’s law and solving the
problem at this iteration step is reduced to solve the usual problem of the theory of
elasticity. As a result, we determine stresses σx1 , σ y1 , σz1 , τx y1 , τ yz1 , τzx1 and strains
7.12 Formula and General Methods 193

εx1 , ε y1 , εz1 , γx y1 , γ yz , γzx1 in the first approximation. Then we determine the stress
intensities σi1 and strain intensities εi1 at each point of the body. On the plane in
coordinates σi − εi (Fig. 7.22), the stress–strain state of some points of the body at
the stage of the first approach is depicted as the point 1 lying on the beam OA. At
the same time tg α ∼ =E∼ = 3G is the angle tangent of beam OA.
In the second approximation, the 3G value must be corrected. In this case
tg α1 = 3G ∗1 = σi1 ∗
/εi1 , where σi1∗
is stress intensity, which corresponds to strain

intensity εi1 , taken from the strain diagram. By values σi1 and εi1 , we obtain param-
∗ ∗ ∗
eters E 1 , μ1 , G 1 that differ at different points in the body. Now, knowing these
parameters, we solve the problem of the theory of elasticity and define stresses
σx2 , σ y2 , σz2 , τx y2 , τ yz2 , τzx2 and strains εx2 , ε y2 , εz2 , γx y2 , γ yz2 , γzx2 in the second
approximation. Then we determine the stress intensities σi2 and strain intensities εi2
at each point of the body. On the plane σi − εi (Fig. 7.22), the stress–strain state of
some points of the body at the stage of the second approach is depicted as the point
2 lying on the beam OB with angle tangent tg α ∼ = 3G ∗1 .
∗ ∗ ∗
In the third approximation tg α2 = 3G 2 = σi2 /εi2 , where σi2 is stress intensity,
which corresponds to strain intensity εi2 , taken from the strain diagram. By values

σi2 and εi2 , we obtain parameters E 2∗ , μ∗2 , G ∗2 . Then we solve the elastic problem
with the specified elastic parameters and determine stresses and strains in the third
approximation. Then we determine the stress intensities σi3 and strain intensities εi3 .
On the plane σi − εi (Fig. 7.22), the stress–strain state of some points of the body at
the stage of the third approach is depicted as the point 3 lying on the beam OC with
angle tangent tg α ∼ = 3G ∗2 and so on (Fig. 7.24).
The calculations should be continued until the results of the calculations of the
nth approximation differ from the results of the calculations of the (n-1)th approxi-
mation by a given value with the required accuracy. Note that the process is always
convergent.

А B С
1 2 3

Fig. 7.24 To the method of elastic parameters


194 7 Mathematical Models of Plasticity Theory

The method of “steps” in the theory of plastic flow


Solving the problem of plastic flow theory presents significant difficulties due to the
fact that the physical equations of plastic flow theory (7.52) contain not only the
components of stresses, but also their increments. Since these equations cannot be
solved with respect to stresses, therefore, it is impossible to construct a system of
equations in displacements.
In many practical problems, numerical integration with the “step-by-step” proce-
dure of the study of the development of plastic deformations is usually used. At each
stage, the external load receives increments from which the corresponding incre-
ments of stresses and strains are calculated. In addition, at each stage, it is necessary
to solve some problems for an elastic-anisotropic body with elastic parameters.
The problem of integrating the equations of plastic flow theory is somewhat simpli-
fied if you can neglect the increments of the elastic strain components compared to
the increments of the plastic strain components.

Control Questions

1. If the yield stress is reached, then yield strains in the direction of the applied
forces:
(1) remain constant;
(2) can reach any value;
(3) disappear.
2. Conditional yield strength:
(1) stress at ε = 0.2%;
(2) stress at ε = εu ;
(3) yield strength at repeated loading.
3. Deformation theory of plasticity connects:
(1) stresses and strains;
(2) stresses and rate of strains;
(3) rate of loading and strains.
4. Theory of plastic flow connects:
(1) stresses and strains;
(2) stresses and rate of strains;
(3) rate of loading and strains.
5. At repeated loading of sample yield strength:
(1) raises;
(2) goes down;
(3) does not change.
7.12 Formula and General Methods 195

6. In case of alternating loading of sample yield stress:


(1) raises;
(2) goes down;
(3) does not change.
7. The theory of perfectly plasticity assumes constancy:
(1) stresses and strains;
(2) strains;
(3) stresses.
8. Tresca-Saint-Venan plasticity condition:
(1) σu = σ y ;
(2) σ1 − σ3 = σ y ;
(3) f 1 (J2d , J3d ) = 0.
9. Huber-Mises-Genky plasticity condition:
(1) σu = σ y ;
(2) σ1 − σ3 = σ y ;
(3) f 1 (J2d , J3d ) = 0.
10. Mises yield surface in axes σ1 , σ2 , σ3 :
(1) cylinder;
(2) prism;
(3) sphere.
11. Saint-Venan yield surface in axes σ1 , σ2 , σ3 :
(1) cylinder;
(2) prism;
(3) sphere.
12. With simple loading, stress tensor components:
(1) proportional to the general parameter;
(2) are extreme;
(3) do not exceed yield strength.
13. Simple loading trajectory in axes σ1 , σ2 , σ3 :
(1) circle;
(2) parabola;
(3) straight line.
14. The dependency of Stress and strain intensity for simple loading is:
(1) linear;
196 7 Mathematical Models of Plasticity Theory

(2) power;
(3) exponential function.
15. In the deformation theory of plasticity, volumetric strain is:

(1) elastic;
(2) plastic;
(3) elastoplastic.

16. In the theory of plastic flow, volumetric strain is:

(1) elastic;
(2) plastic;
(3) elastoplastic.

17. In the deformation theory of plasticity, stress and strain deviators are:
(1) proportional;
(2) linearly dependent;
(3) not linked.

18. In the theory of plastic flow, the stress deviator is connected to:

(1) strain deviator;


(2) deviator of plastic strains;
(3) rate of strains.

19. In the deformation theory of plastic material is:

(1) not hardening;


(2) hardening;
(3) perfectly plastic.

20. Hardening hypothesis of the deformation theory of plasticity:

(1) σu = .(ε (∫u );p )


(2) σu = . εu ;
(3) si j = 23 σξuu ξi j .

21. Hardening hypothesis of the theory of plastic flow:

(1) σu = .(ε (∫u );p )


(2) σu = . εu ;
(3) si j = 23 σξuu ξi j .

22. In the method of elastic solutions at the first step is zero:

(1) strain rate;


(2) volumetric strain;
(3) plasticity function.
7.12 Formula and General Methods 197

23. With increasing of the strain rate, the yield strength:

(1) does not change;


(2) increases;
(3) decreases.

24. Volterra principle: the solution of the viscoelastic problem can be obtained from
the solution of the corresponding problem:

(1) elasticity;
(2) plasticity;
(3) fracture theory.

25. Draker’s postulate refers to criteria:

(1) hardening;
(2) softening;
(3) perfectly plasticity of the material.

26. Gradient hypothesis: strain increment vector is directed to the yield surface:

(1) along the tangent;


(2) normal;
(3) at an angle 45° .

27. The flow surface separates the elastic region from the:

(1) plastic area;


(2) viscoelastic area;
(3) destruction area.
Chapter 8
Fundamental Solutions

8.1 General Remarks

Fundamental solutions relate to solve boundary problems in case of point loading


action. This is a finite force applied at a point: a surface of zero area. The point load is
a mathematical artifice because we cannot achieve it in reality. Intuitively, we expect
one or more components of stress to become infinite at the point where the load
is placed, and this is realized in the mathematical solutions. Because of the stress
singularities, understanding point-load problems will involve limiting procedures,
which, as we know, are a bit dubious in regards to solids. But this should not deter
us, since we will show the applications where the stresses remain finite and the
applied loads are not point loads.
When constructing mathematical models of solid mechanics problems, in most
cases, we are dealing with the solution of boundary problems for elliptic equations.
Therefore, it seems logical to consider primarily fundamental solutions for elliptic
equations.
Fundamental solutions to elliptic equations play an important role in constructing
the general theory of elliptic boundary problems. Therefore, for example, they allow
you to obtain integral representations of solutions, investigate their properties and
reduce the boundary problem to integral equations.
A classic example of an elliptical equation is the Laplace equation:

Σ
n
∂ 2u
Δ x u ≡ = 0, n = 2, 3. (8.1)
i=1
∂ xi2

The fundamental solution of the Laplace Eq. (8.1) is a function in the following
form:
[
⎧ 1 | n
− 4πr , n = 3, |Σ
K 1 (x, y) = 1 r =] (xi − yi )2 .

ln 1
r
, n = 2,
i=1

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 199
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_8
200 8 Fundamental Solutions

The function K 1 (x, y) is a solution to such differential equation:

Δ x K 1 (x, y) = δ(x, y), (8.2)

where δ (x − y) is the Dirac1 function.


Solution (8.2) has the property that for any infinitely differentiable function ϕ(x)
with a compact carrier, the following identity takes place:
∫ ∫ ∫ ∫
ϕ(x) = (n) K 1 (x, y)Δ y ϕ(y)dy = Δ x (n) K 1 (x, y)ϕ(y)dy.

or
∫ ∫
(n) ϕ(y)δ(x − y)dy = ϕ(x).

Denote via L differential operator:

Σ
n Σ
2m
∂ i1 +...+ik u
Lu = ai1 ...ik . (8.3)
k=0 i 1 +...+i k =1 ∂ x1i1 . . . ∂ xkik

The fundamental solution K (x, y) of the operator (8.3) is determined by the


following equation:

Lx K (x, y) = δ (x − y). (8.4)

Formula (8.4) should be understood as follows:


∫ ∫
L (n) K (x, y)ϕ(y)dy = ϕ(x).
...

Note that for second-order elliptic equations with constant coefficients in the
following form:

Σ
n
∂ 2u
ai j = 0, ai j = const, i, j = 1, n (8.5)
i, j=1
∂ xi ∂ x j

the fundamental solution is obtained by converting the corresponding quadratic form


to a canonical form. As a result:

1 Paul Adrien Maurice Dirac (1902–1984), British theoretical physicist who is regarded as one of
the most significant physicists of the twentieth century.
8.2 Boussinesq’s Problem 201
⎧ ( ) 2−n

⎪ Σ n ( ) 2

⎪ Ai j (xi − yi ) x j − y j ,n>2
⎨c
i, j = 1
K (x, y) = ( )

⎪ Σ
n ( )

⎪ Ai j (xi − yi ) x j − y j , n = 2,
⎩ c ln
i, j = 1

|| ||n
where Aij is the algebraic complement to aij in matrix ||ai j ||i, j=1 ; c is a constant
defined by coefficients (8.5).

8.2 Boussinesq’s Problem

Of all the point load problems, the most useful in solid mechanics is the problem of
a point acting normal to the surface of an elastic half-space (Fig. 8.1).
The Boussinesq’s problem is to determine stress–strain state in an elastic isotropic
and homogeneous half-space x3 ≥ 0, loaded at the origin of coordinates by a point-
load directed along the x 3 axis. It has magnitude P: p3 (x1 , x2 ) = Pδ(x1 )δ(x2 ).
Because the load is symmetrical about the planes (x 1 x 3 ) and (x 2 x 3 ), cylindrical
coordinates are most convenient here (Fig. 8.1).
The boundary conditions for this problem are as follows. Everywhere on the
surface x3 = 0, except at the origin (0,0,0) or r = 0, loads are specified to zero. At
the origin, the stresses equilibrate the applied load P, that is:

p1 = p2 = 0, p3 (x1 , x2 ) = Pδ(x1 )δ(x2 ) = −σ33 (x1 , x2 , 0), x3 = 0.

Besides, for any point in the half-space infinitely distant form the origin, all the
displacements must vanish:
.
x) → 0, x → ∞, i = 1, 2, 3.
u i (→

For solution building of this problem, we can introduce such guess: the solution must
have radial symmetry. That means nothing can depend on the θ coordinate, that is,

Fig. 8.1 Boussinesq’s P


problem

x1
θ

x2 ψ a
x3
202 8 Fundamental Solutions

whatever the angle θ we choose, the solution must be exactly the same. The second
guess is that the component of the displacement u θ must be zero. If u θ were not zero
everywhere, there would be torsion of the half-space, at least, at some points (as the
point load could not cause any torsional motions). Therefore, displacement vector
has the form:

u→ = (u r , 0, u z ),

where u r and u z are not functions of the angle θ.


Now to get expressions for strains, we will use representation for strains in the
form:
1( )
e= ∇ u→ + (∇ u)
→ T , (8.6)
2

where e is the strain matrix, ∇ u→ is the displacement gradient matrix. In cylindrical


coordinates, the displacement gradient matrix looks like this:
⎡ ⎤
∂u r ∂u r u θ ∂u r
∂r
1
∂θ
− r ∂z
⎢ ∂u θ
r
∂u θ u r ∂u θ ⎥
∇ u→ = ⎣ ∂r
1
r ∂θ
+ r ∂z ⎦.
∂u z 1 ∂u z ∂u z
∂r r ∂θ ∂z

If u θ is zero, u r and u z do not depend upon θ, then we see that ∇ u→ will have
only five non-zero components lying on the two diagonals of the matrix. The strain–
displacement relationship (8.6) tells us that there will similarly be only five non-zero
strain components. Finally, if we use Hooke’s law, we see the stress matrix must have
this form:
⎡ ⎤
σrr 0 σr z
σ = ⎣ 0 σθθ 0 ⎦,
σr z 0 σzz

where the stress σzz is the axial stress; σrr is the radial stress; σθθ is the hoop stress.
Only one non-zero shear stress is present.
The procedures for building up the Boussinesq’s solution are described in various
textbooks. Here we just write out Boussinesq’s solution for displacements and stress
components for the vertical point-load problem.
In Cartesian coordinates:
[ 2 ]
P x3 1
u3 = + 2(1 − ν) ,
4πμ R 3 R
[ ]
P · xi x3 1
ui = − (1 − 2ν) , i = 1, 2. (8.7)
4πμ R 3 R(R + x3 )
8.2 Boussinesq’s Problem 203

3P x j x32
σ3 j = − · , j = 1, 3
2π R 5
[ ( 2 )]
3P x12 x3 1 R + R · x3 + x32 x12 (2R + x3 )
σ11 =− + (1 − 2ν) − ,
2π R 5 3 R 3 (R + x3 ) R 3 (R + x3 )2
[ ( 2 )]
3P x22 x3 1 R + R · x3 + x32 x22 (2R + x3 )
σ22 =− + (1 − 2ν) − ,
2π R 5 3 R 3 (R + x3 ) R 3 (R + x3 )2
[ ]
3P x1 x2 x3 1 x1 x2 2R + x3
σ12 =− − (1 − 2ν) . (8.8)
2π R5 3 R 3 (R + x3 )2

In cylindrical coordinates:
[ ]
P rz r
ur = − (1 − 2ν)
4πμR R 3 (R + z)
uθ = 0
[ 2 ]
P z
uz = + 2(1 − ν) (8.9)
4πμR R 2
[ ]
P 3r 2 z (1 − 2ν)
σrr = − − ,
2π R 5 R(R + z)
[ ]
P(1 − 2ν) z 1
σθθ = − − 3+ ,
2π R R(R + z)
[ 3]
P 3z
σzz = − ,
2π R 5
[ ]
P 3r z 2
σr z = σzr = − ,
2π R 5
σr θ = σθr = σθz = σzθ = 0. (8.10)

In these equations R 2 = r 2 + z 2 . For any point in the half-space, R is the straight-


line distance to the origin.
Considering the stress field in Eqs. (8.10), we note that as R becomes larger, all
the stress components approach to zero. On the boundary z = 0, we see that σzz
and σr z vanish at every point, except at the origin of coordinates. At the origin, the
stresses become singular, just as we expect for a point load. The unit normal vector
to the boundary is n→ = (0, 0, −1). If we use this with equation for traction vector
T = σT n,→ we find that the traction vector on the boundary is zero everywhere except
at origin. We’ll see how the singular stresses exactly equilibrate the point load in a
moment.
Next, let’s consider the displacement field. If we let R become larger, both u r and
u z approach to zero, as our boundary conditions specified. On the boundary x 3 = 0
(z = 0), we have:
204 8 Fundamental Solutions

P · xi (1 − 2ν) (1 − ν)P ( )1 2
ui = − 2
, i = 1, 2, u 3 = , r = x12 + x22 / ;
4πμr 2πμr
P(1 − 2ν) P(1 − ν)
ur = − , u zz = .
4πμr 2πμr

These formulas show that displacements become singular at the origin.


Now let’s find out how the stress field equilibrates the point load. For this purpose,
we consider the hemispherical surface of radius a (Fig. 8.1). Let ψ be the angle
between a radius of the hemisphere and the z axis (Fig. 8.1). For any point on this
surface, we have R = a = constant. The unit normal vector of the surface at any point
can be written as:
⎛ ⎞
sin ψ
n→ = ⎝ 0 ⎠. (8.11)
cos ψ

Components of the point r and z are:

z = a cos ψ, r = a sin ψ. (8.12)

The components of stress at any point on the surface can be obtained by setting
R = a and using (8.12) and (8.10):
[ ]
P 3 sin2 ψ cos ψ (1 − 2ν)
σrr = − − ,
2π a2 a 2 (1 + cos ψ)
[ ]
P(1 − 2ν) cos ψ 1
σθθ =− − 2 + 2 ,
2π a a (1 + cos ψ)
3P cos3 ψ
σzz =− ,
2πa 2
3P sin ψ cos2 ψ
σr z = σzr = − ,
2πa 2
σr θ = σθr = σθz = σzθ = 0. (8.13)

Using (8.11), we can obtain the traction vector that acts on the hemispherical
surface:
⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
Tr σrr 0 σr z sin ψ σrr sin ψ + σr z cos ψ
T = ⎣ Tθ ⎦ = ⎣ 0 σθθ 0 ⎦⎣ 0 ⎦ = ⎣ 0 ⎦. (8.14)
Tz σr z 0 σzz cos ψ σr z sin ψ + σzz cos ψ

And, finally, we use (8.13) to completely specify T in terms of P, a, and ψ.


Now some words about the overall equilibrium of that part of the half-space
contained inside the hemispherical surface. The horizontal upper surface just supports
8.2 Boussinesq’s Problem 205

Fig. 8.2 Geometry for P


integrating to find resultant
force on hemispherical
surface
r
a
ψ

adψ

the vertical force P. The hemispherical surface is acted on the tractions given in
Eq. (8.14). These tractions have one horizontal component T r and one vertical compo-
nent T z . The horizontal components exactly cancel. For every T r there is an equal and
opposite component on the opposing side of the hemisphere. The vertical compo-
nents T z do not cancel. They have to equilibrate the applied load P. We can write out
T z by using the equations of (8.13) and (8.14):

3P [ 2 ] 3P
Tz = − 2
sin ψ cos2 ψ + cos4 ψ = − cos2 ψ. (8.15)
2πa 2πa 2
If we integrate T z over the hemispherical surface, we will find the resultant of
upward forces, which equal to P. Let’s show it. Considering the elementary site
around the hemispherical surface (Fig. 8.2) with the width a dψ, the total length of
the elementary site is 2πa sin ψ.
For calculating the total upward force acting on the elementary site we use (8.15):

Tz Sar ea = Tz 2πa 2 sin ψdψ = 3P sin ψ cos2 ψdψ, (8.16)

where Sar ea is total area of the elementary site.


From (8.16), we see that the upward force on the site depends upon ψ and P, but
not on a, that is, if we hold ψ and P constant, then this site force will be the same
on every hemispherical surface no matter how big or how small the size of a is. And
finally, we can find the resultant of the upward force F s that acts on the hemispherical
surface by integrating:

∫π/ 2
Fs = 3P sin ψ cos2 ψdψ = P.
0

F s exactly equals to P, which is what we expected to obtain. We emphasize again


that this conclusion is true for any radius a. Even if we let a approach to zero, and
hence have infinite stresses, the resultant force will exactly equilibrate P. Thus, we
206 8 Fundamental Solutions

have found that all the specified boundary conditions are satisfied. Now we verify that
the equilibrium equations are also satisfied. For cylindrical coordinates, equilibrium
equations have the form:

∂σrr 1 ∂σr θ ∂σr z 1


+ + + (σrr − σθθ ) + fr = 0,
∂r r ∂θ ∂z r
∂σrθ 1 ∂σθθ ∂σθz 2
+ + + σr θ + f θ = 0,
∂r r ∂θ ∂z r
∂σr z 1 ∂σθz ∂σzz 1
+ + + σr z + f z = 0. (8.17)
∂r r ∂θ ∂z r

If we substitute expressions for the stresses (8.10), we will find these Eqs. (8.17)
are exactly satisfied for the case where the body forces ( fr , f θ , f z ) are zeros.
Remark Boussinesq’s solution has a lot of applications, especially in geotechnical
engineering.

8.3 Hertz’s Formulas

Consider the following problem: in the plane x 3 = 0, which bounds the elastic
half-space, within the region S p there is a normal load p3 = p(x1 , x2 ) (Fig. 8.3).
For the correctness
˜ of the setting of the boundary problem, we will assume that
the integral | p(x1 , x2 )|d x1 d x2 is the limited value.
To find the stress–strain state caused by such a load, we use the Boussinesq’s
solution for the point load acting at the point x = (x1 , x2 , 0).
We will use the formulas (8.7), rewriting them in the following form:

Fig. 8.3 Principle loading


x1 р3
pattern for Hertz’s problem

x2

x3
8.3 Hertz’s Formulas 207

( − [ ]
→) (1 − ν)Pδi3 P ∂ x3
→ ξ =
Ui x, − · + (1 − 2ν) ln(R(x, ξ) + x3 ) ,
π μ R(x, ξ) 4π μ ∂ xi R(x, ξ)
(8.18)
( )1 2
where R(x, ξ) = (x1 − ξ1 )2 + (x2 − ξ2 )2 + x32 / is the distance between points
x and ξ.
Let’s integrate the expression (8.18) over the area S p in the plane x3 = 0 on which
the distributed load acts:
¨
(1 − ν)δi3 p(ξ1 , ξ2 )
x) =
u i (→ dξ1 dξ2
πμ R(x, ξ)
Sp
¨ [ ]
1 ∂ x3
− · p(ξ1 , ξ2 ) + (1 − 2ν) ln(R(x, ξ) + x3 ) dξ1 dξ.
4π μ ∂ xi R(x, ξ)
Sp
(8.19)

Enter the following functions:


¨
.(→
x) = p(ξ1 , ξ2 ) ln[R(x, ξ) + x3 ]dξ1 dξ2 ,
Sp
¨
p(ξ1 , ξ2 ) ∂ .(→
x)
.(→
x) = dξ1 dξ2 = . (8.20)
R(x, ξ) ∂ x3
Sp

From the definition of the function .(→ x) (8.20), it follows that it satisfies the
Laplace equation outside the region S p , it is a continuous function everywhere except
for the region S p , and tends to zero at R → ∞. The derivative of the function .(→ x)
in the direction of normal to the boundary of the half-plane has a break: when moving
from point x→ = (x1 , x2 , x3 ) to point x→ 0 = (x1 , x2 , 0) on the side of positive values of
x 3 , the following relation occurs:
| ⎧
∂ . || −2π p(x1 , x2 ), x ∈ S p ,
| = (8.21)
∂ x3 x3 →+0 0, x∈/ Sp.

The function .(x) in half-space is harmonic. At R → ∞ this function increases


indefinitely. Derivatives of .(x) by coordinates in the half-space x3 > 0 at R → ∞
tend to zero.
Using the functions .(→
x) and .(x), relation (8.19) can be represented in the
following form:

1 − ν → 1
u→ = i3 . − grad[x3 . + (1 − 2ν).]. (8.22)
π·μ 4π μ
208 8 Fundamental Solutions

Rewrite (8.22) in the components of the displacement vector:


[ ] [ ]
1 ∂. ∂. 1 ∂. ∂.
u1 = − x3 + (1 − 2ν) , u2 = − x3 + (1 − 2ν) ,
4π μ ∂ x1 ∂ x1 4π μ ∂ x2 ∂ x2
1−ν 1 ∂.
u3 = .− x3 . (8.23)
2π · μ 4π μ ∂ x3

Based on (8.23), the displacement values on the plane x3 = 0 are defined by the
following formulas:

1 − 2ν ∂.(x1 , x2 , 0) 1−ν
uβ = − , β = 1, 2 u 3 = .(x1 , x2 , 0).
4π μ ∂ xβ 2π · μ

For the stress components, we get:


( )
x3 ∂ 2 . x3 ∂ 2 . 1 ∂. ∂ 2.
σ13 =− , σ23 = − , σ33 = − x3 2 . (8.24)
2π ∂ x1 ∂ x3 2π ∂ x2 ∂ x3 2π ∂ x3 ∂ x3

From (8.24), it is easy to obtain that at x 3 = 0 it is performed that:


( )
1 ∂.
σ31 = 0, σ32 = 0, σ33 = .
2π ∂ x3 x3 →+0

From (8.21), we obtain that:


( )| ⎧
1 ∂ . || − p(x1 , x2 ), x ∈ S p ,
σ33 (x1 , x2 , 0) == =
2π ∂ x3 |x3 →0 0, x∈/ Sp.

If you enter a new function:

1
F =− [x3 . + (1 − 2ν).],
4π μ

then the displacements (8.23) can be represented by a shorter record, namely:

∂F 1−ν
ui = + δi3 .. (8.25)
∂ xi πμ

Formulas (8.25) were first introduced by Hertz and were used by him in solving
contact problems.

Remarks The problems connected with determination of the stress–strain state in


an elastic half-space in case of action on boundary x 3 = 0 loadings, perpendicular to
the plane x 3 = 0, were solved by many researchers.
8.4 Flamant’s Problem 209

Fig. 8.4 Flamant’s problem


for a half-space p
dy

xa

ba
y Ra
za
x
A
φ
z

8.4 Flamant’s2 Problem

8.4.1 Flamant’s Problem for a Half-Space

In this section, we consider a line load acting on the surface of a half-space (Fig. 8.4).
The line load acts on an infinitely long line. The intensity of this line load is p, and
it has dimensions of force per unit length.
We will use the Boussinesq’s solution together with the principle of superposition
to solve the stress–strain state in the half-space.
Consider the point A in Fig. 8.4. Note, the fact that this point lies beneath the x
axis doesn’t mean any loss of generality. Since the load extends to infinity in both
directions on the y axis, we can place the origin of coordinates wherever we like on
the y axis.
Let’s find the σzz components of stress at point A due to the action of the line load
on the surface of a half-space. We consider an elementary length dy of the y axis
(Fig. 8.4). For any point inside the half-space, including point A, the line load p acting
on the element dy may be considered as a point load. There will be an increment of
stress dσzz caused by this “point load” pdy. According to the Boussinesq’s solution
(8.10), we have:
[ ]
( pdy) 3z 3
dσzz = − . (8.26)
2π R5

The distances between z and R are shown in Fig. 8.4.


To obtain the stress σzz , we must integrate both sides of (8.26):

∫∞
3 pz 3
σzz = − dy. (8.27)
2πR 5
−∞

2 Flamant A. (1839–1915), French mechanical engineer.


210 8 Fundamental Solutions

( )1 2
Note that (Fig. 8.4) b = x 2 + z 2 / , y = btgϕ. So dy = b sec2 ϕdϕ. We can
rewrite (8.27) as:

∫π/ 2
3 pz 3
σzz = − cos3 ϕdϕ (8.28)
2πb4
−π/ 2

Note that z and b in (8.28) are constants. Now we integrate (8.28) and obtain:

2 pz 3 2 pz 3
σzz = − = − ( )2 . (8.29)
πb4 π x 2 + z2

The other components of stress are obtained by similar methods:

2 px 2 z
σx x = − ( )2 ,
π x 2 + z2
2 pyz
σ yy = − ( 2 ),
π x + z2
2 px z 2
σx z = − ( )2 = σzx ,
π x 2 + z2
σx y = σ yx = σzy = σ yz = 0. (8.30)

Remark Obtaining all these components is not quite easy as obtaining σzz .
The vertical stress component σzz doesn’t differentiate between cylindrical and
rectangular coordinates, but the other stress components do, and we must be careful.

Note that when we integrated (8.26) to get (8.27), we were using the principle of
superposition. However, if we did not deal with linear elasticity, this might not be
possible.
Let’s consider a cylindrical surface of radius b whose axis coincides with direction
of line load (Fig. 8.5).
Let’s obtain the tractions that act on this surface by using the expressions for
stresses (8.29) and (8.30). As a result of performing mathematical operations, we
get:

2 pz
T=− →
n, (8.31)
πb2

where n→ is the unit normal to the cylindrical surface.


If we integrate the vertical component of the traction (8.31) around the cylindrical
surface, we can make sure that the resultant of upward forces exactly equilibrates
the applied load p.
8.4 Flamant’s Problem 211

Fig. 8.5 Cylindrical surface p


whose axis coincides with
direction of line load

x
b
α

From (8.31), we can conclude that the cylindrical surface itself is a principal
surface. The minor principal stress acts on it:

2 pz
σ1 = − . (8.32)
πb2

Remark The principal stress σ1 is the major by absolute value (by modulus).

The intermediate principal surface is defined by n→ = (0, 1, 0) and the interme-


diate principal stress is σ2 = νσ 1 . The major principal surface is perpendicular to
the cylindrical surface and to the intermediate principal surface. The major principal
stress is exactly zero.
Some words about the distribution of the principal stress in space. Consider the
locus of points on which the major principal stress σ1 is a constant. From (8.32), we
see this will be a surface for which:
z π|σ1 | 1
= = ,
b2 2p 2c

where c is a constant. From this ratio, it follows that:

b2 = x 2 + z 2 = 2cz.

This is the equation of a circle with the radius c centered on the z axis at a depth c
beneath the origin. At each point on this circle, the major principal stress is the same.
It points directly to the origin. If we were to consider a larger c and consequently a
larger circle, the value of σ1 would be smaller in inverse proportion to c. This result
allows us to formulate the idea that there is a pressure bubble under the foundation
in the soil.
212 8 Fundamental Solutions

8.4.2 Flamant’s Problem for a Half-Plane

Interesting point is when we are dealing with a plane strain problem. For this type of
problem e yy = ex y = e yx = ezy = e yz = 0. Also, the non-zero strains are not func-
tions of y. A particle that initially has coordinate y0 in the referred configuration will
always have coordinate y0 in any deformed configuration unless a rigid translation
in the y direction occurs.
We will make some comments regarding the selection of a coordinate system
when constructing model problems in a two-dimensional statement. This is important
since the displacements of the points of the elastic body during its deformation are
determined before additive displacements occur.

Remarks Additive displacements are displacements of a deformed body with respect


to some selected coordinate system as a rigid body.

Since the components of strains and stresses in the elastic half-plane decrease,
/ as
they are removed from the place of application of the load, with order 1 r , when
they are integrated (when determining displacements), logarithmic features appear.
Therefore, with a poorly selected coordinate system, when moving away from the
point of application of the load (r → ∞), displacements may be infinite. Note that
in the case of considering a three-dimensional elastic body, the decrease in strains
/ 2
and/stresses is in order 1 r . Accordingly, for displacements, the decreasing order
is 1 r .
Some explanations in this regard are given as below.
Consider an elastic half-space, to the boundary of which an arbitrary load is applied
on a bounded section with a finite resultant force and a finite principal moment.
According to the above facts, the origin of the coordinate system can be associated
with a half-space point, as far as you want from its boundary. One of the axes can
be directed along an element that was perpendicular to the half-space boundary
before deformation. The displacements of the elastic half-space points in such a
coordinate system are always finite and disappear as they move away from the place
of the boundary where the load is applied. As a result, this coordinate system can
be replaced by another coordinate system, starting at one of the points of the body
boundary before deformations, with two axes lying in the boundary plane and a third
axis perpendicular to it. Displacements of a half-infinite solid with respect to this
new coordinate system will be simultaneous displacements with respect to infinitely
distant elements. That is, you can “fix” the body at infinitely distant points.
Unfortunately, a similar procedure cannot be done in the case of an elastic half-
plane, even if a load is applied at a section of the boundary of the finite length.
Indeed, in this case, the change in distance (and therefore displacements) between two
arbitrary points of a 2D body (located, for example, perpendicular to the undeformed
half-plane boundary) grows according to the logarithm law depending on the distance
itself. Therefore, if you take a point at infinity as the origin (i.e., “fix” half-plane in
infinity), then the displacements of boundary are unlimited. That is, the origin of the
coordinate system with respect to which the displacements are determined must be
8.4 Flamant’s Problem 213

a b x

Fig. 8.6 Flamant’s problem for a half-plane

associated with a particular point of the elastic half-plane (and not be arbitrary, as in
the case of half-space). Thus, it can be, for example, a point located on the boundary
of the half-plane.
A. Consider the problem of the action of a point load P→ on an elastic half-plane
(Fig. 8.6). The load P→ is directed along the z axis and applied at the origin.
In order to correctly set the problem, as mentioned above, it is necessary to
clearly indicate the coordinate system with respect to which the displacements
are calculated. We fix the half-plane element located on the z axis at the depth
h. In this case, the following equalities are fulfilled: at x = 0, z = h: u = 0,
v = 0, ∂z u = 0, where u = u(x, z) and v = v(x, z) are correspondingly the
components of the displacement of the points of the elastic half-plane parallel
to the axes x and z.
The solution of the formulated problem on the effect of the point load on the
elastic half-plane, taking into account the fixing conditions, can be presented in the
following form:
[ ] [ ]
ϑ λ+μ xz 1 r λ + μ x2
u=A − , v = A ln + ,
2(λ + 2μ) 2μ(λ + 2μ) r 2 2μ h 2μ(λ + 2μ) r 2
λ + 2μ P .
A=− , r = x 2 + z 2 , ϑ = arctg(x / z). (8.33)
λ+μ π

For points of elastic half-plane boundary, formulas (8.33) take the following form (z
= 0):

A π
u(x, 0) = U (x) = sign x,
2(λ + 2μ) 2
[ ]
1 |x| λ+μ
v(x, 0) = V (x) = A ln + . (8.34)
2μ h 2μ(λ + 2μ)
214 8 Fundamental Solutions

Fig. 8.7 To problem about dξ


the normally distributed load
p(ξ) x
acts normally on horizontal
boundary of elastic
half-plane
ξ h

B. Now let’s consider the next problem. The normally distributed load of intensity
p(ξ) acts normally on horizontal boundary of elastic half-plane in section S :
ξ ∈ [a, b] (Fig. 8.6). Define the displacements of half-plane boundary points.

If a load pdξ acts at the origin, then the displacements on the boundary element,
remote from the point (0, 0), by the distance ξ (Fig. 8.7), in accordance with the
formula (8.34), are:

λ + 2μ |ξ|
dVf = − ln pdξ. (8.35)
2 π μ(λ + μ) h

If the load p(ξ) is distributed in the section ξ ∈ [a, b], then the displacements at
the origin are:

∫b
λ + 2μ |ξ|
V f (x) = − ln p(ξ)dξ. (8.36)
2 π μ(λ + μ) h
a

Formulas (8.35–8.36) for determining displacements at the origin are not accurate.
Therefore, the expression (8.35) for displacements at the origin from the load pdξ is
true in the assumption of fixing the half-plane element at the depth h, located directly
below the point of application of the load, that is, at the distance ξ from the z axis
(Fig. 8.7). Thus, in the expression (8.35), in fact, summation of displacements of an
infinite number of deformed states with fixations at different points ξ is performed,
which is fundamentally unacceptable.
We describe the procedure for constructing the correct solution
A. Let the point load act at the points of the boundary with the coordinate (ξ, 0).
Formulas for displacements are obtained from formulas (8.33), in which the
coordinate x is replaced by the difference (x − ξ). In addition, displacements
corresponding to the additive displacement of the half-plane as a rigid body
should be added to the values u and v. As a result, we get:
8.4 Flamant’s Problem 215

A x −ξ λ+μ (x − ξ)z
u= arctg − A 2 + α + ω z,
2(λ + 2μ) z 2μ(λ + 2μ) z + (x − ξ)2
/
A λ+μ (x − ξ)2
v= ln z 2 + (x − ξ)2 + A 2 + β − ω x.
2μ 2μ(λ + 2μ) z + (x − ξ)2
(8.37)

To determine the additive displacement constants α, β, ω, we assume that at


the point (0, h), the displacements u and v are turned to zero along with the deriva-
tive ∂ u/∂ z (thereby fixing the element of the axis z at the depth h). Solving the
corresponding equations, we get:

A ξ A h·ξ λ+μ 2h 3 ξ
α= arctg + − ,
2(λ + 2μ) h 2(λ + 2μ) h 2 + ξ2 2μ(λ + 2μ) (h 2 + ξ2 )2
A . 2 λ+μ ξ2
β=− ln h + ξ2 − A 2 ,
2μ 2μ(λ + 2μ) h + ξ2
( 2 )
A ξ λ+μ ξ − h2 ξ
ω=− − A( ) .
2(λ + 2μ) h 2 + ξ2 2μ(λ + 2μ) h 2 + ξ2 2

We substitute these values α, β, ω into expressions (8.37) for u and v, and after
obvious transformations, we have:
[ ]
A x −ξ ξ hξ yξ
u(x, z, ξ) = arctg + arctg + 2 −
2(λ + 2μ) z h h + ξ2 h 2 + ξ2
[ ( 2 ) ]
λ+μ (x − ξ)z 2h 3 ξ ξ − h2 ξ z
− A 2 +( )2 + ,
2μ(λ + 2μ) z + (x − ξ)2 h 2 + ξ2 h 2 + ξ2
/
A (x − ξ)2 + z 2 A ξx
v(x, z, ξ) = ln +
2μ h +ξ
2 2 2(λ + 2μ) h + ξ2
2
[ ( 2 ) ]
λ+μ (x − ξ)2 ξ2 ξ − h2 ξ x
+ A 2 − 2 + ( )2 .
2μ(λ + 2μ) z + (x − ξ)2 h + ξ2 h 2 + ξ2
(8.38)

The formula for determining the vertical displacements of the boundary surface
points (at z = 0), on the basis of (8.38), is as follows:

A |x − ξ| A ξ· x
V (x, ξ) = v(x, 0, ξ) = ln . +
2μ h +ξ
2 2 2(λ + 2μ) h + ξ2
2
[ ( 2 ) ]
λ+μ ξ2 ξ − h2 ξ · x
+ A 1− 2 + ( )2 (8.39)
2μ(λ + 2μ) h + ξ2 h 2 + ξ2
216 8 Fundamental Solutions

As a result, (8.39) is the displacement on the surface of the elastic half-plane as


a result of the action at the point (ξ, 0) of the point load P→ perpendicular to the
boundary of the elastic half-plane.
B. Now let’s look at the case where the distributed load p(ξ) is acting on the site
(a, b). Vertical displacements from such load at any point (x, 0) of the boundary
according to (8.39) in this case have the following representation:

∫b ⎧
1 λ + 2μ 1 |x − ξ| 1 ξ· x
V (x) = p(ξ) ln . +
π λ+μ 2μ h +ξ
2 2 2(λ + 2μ) h + ξ2
2
a
[ ( 2 ) ]⎧
λ+μ ξ2 ξ − h2 ξ · x
+ A 1− 2 + ( )2 dξ. (8.40)
2μ(λ + 2μ) h + ξ2 h 2 + ξ2

Compare the obtained formulas (8.40) with inaccurate expressions (8.36). Let
x = 0, a = 0, b = c, λ = μ, p = p0 = const.
According to the exact formula (8.40), we obtain:

∫c ⎧ . ⎧
3 · p0 2 2 ξ2
V (0) = ln h + ξ − ln ξ − +
2 2 dξ.
4π μ 3 3 h 2 + ξ2
0

According to the imprecise formula (8.36), we have:

∫c
3 · p0
V f (0) = (ln h − ln ξ)dξ.
4πμ
0

After calculating the integrals, we finally get:


( √ )
3 · p0 h 2 + c2 2 c
V (0) = c ln + h arctg ,
4π μ c 3 h
( )
3 · p0 h
V f (0) = c ln + c .
4πμ c

It follows that when using the formula for V f , we get the wrong result, if we take
in it that h = 0 (i.e., fix the half-plane element to the origin (0, 0)). In this case, the
displacements become infinitely larger. The exact formula for V at (0, 0) gives the
value V (0) = 0.
8.5 Kelvin’s Problem 217

8.5 Kelvin’s Problem

In this section, we consider another point load problem. This is the problem of a
point load acting in the interior of an infinite elastic body (Fig. 8.8).
The procedure for building up the Kelvin’s solution base on Kelvin’s methods
and it is described in detail, for example, in textbooks [Nowacki, W. Theory elas-
ticity/W. Nowacki.—Moscow. Pub. by “Mir”, 1975. 872 p. (in Russian), Zhuravkov,
M.A., Continua Mechanics. Theory of Elasticity and Plasticity. Coursebook/M.A.
Zhuravkov, E.I. Starovoitov. Minsk. Pub. by Belarusian State University, 2011. 543
p. (Classical University Edition). (in Russian)].
The Kelvin’s solution, written in cylindrical coordinates has such form:

Pzr
ur = ,
16πμ(1 − ν)R 3
u θ = 0,
[ 2 ]
P z 1 2(1 − ν)
uz = + + . (8.41)
16πμ(1 − ν) R 3 R R
[ 2 ]
P 3r z (1 − 2ν)z
σrr = − − ,
8π(1 − ν) R 5 R3
P(1 − 2ν)z
σθθ = ,
8π(1 − ν)R 3
[ 3 ]
P 3z (1 − 2ν)r
σzz = − + ,
8π(1 − ν) R 5 R3
[ ]
P 3r z 2 (1 − 2ν)r
σr z = σzr = − + ,
8π(1 − ν) R 5 R3

Fig. 8.8 Kelvin’s problem

x1
θ

x2
x3/z
218 8 Fundamental Solutions

Fig. 8.9 Geometry for P


integrating vertical stress σzz

ψ R r
c

r dr

σr θ = σθr = σθz = σzθ = 0, (8.42)

where R 2 = z 2 + r 2 .
Solutions (8.41–8.42) have singularities at the origin, where the point load acts,
and both displacements and stresses die out for large R. Note that on the plane z = 0,
all the stress components except for σr z vanish at all points except for the origin.
Now we investigate the solutions (8.41–8.42) by considering a planar surface
defined by z = c. The vertical component of traction on this surface is σzz . Let’s
now integrate σzz over this entire surface for calculating the resultant force. Note that
the value of σzz will be a constant on any horizontal circle centered on the z axis.
Therefore, the force acting on the annulus (Fig. 8.9) will equal to σzz multiplied by
the area 2πr dr . The total resultant force Rsum on the surface z = c is given by the
following formula:

∫∞ ∫∞ [ 3 ]
P 3c (1 − 2ν)c
Rsum = σzz 2πr dr = − + r dr .
4(1 − ν) R 5 R3
0 0

For integration, we introduce the angle ψ (Fig. 8.9). We have: r = ctgψ and
dr = c sec2 ψdψ, taking into account these relations, the integral takes the form:

∫π/ 2
P [ ]
Rsum = − (1 − 2ν) sin ψ + 3 cos2 ψ sin ψ dψ.
4(1 − ν)
0

After integral calculation, we find that the resultant force equals to –P/2, exactly
one-half of the applied load. If now we consider a similar surface z = −c, we find
tensile stresses of the same magnitude as the compressive stresses on the lower
plane. The resultant force on the upper plane is P/2 (a tensile force). Together, the
two resultant forces exactly equilibrate the applied load.
8.5 Kelvin’s Problem 219

As we noted earlier, on the plane z = 0, all the stress components except for
σr z vanish
/ at all points except for the origin. Let’s consider the special case, where
ν = 1 2 (an incompressible material). Then σr z will also be zero on this surface, and
that part of body below the z = 0 plane becomes equivalent to the half-space of the
Boussinesq’s problem.
/ Comparing the Kelvin’s solution with Boussinesq’s solution
for the case ν = 1 2, we see they are identical for all z ≥ 0. For z ≤ 0 we also have
the Boussinesq’s solution, but with a negative load –P. The two half-spaces, which
together comprise the infinite body of Kelvin’s problem, act as if they are uncoupled
on the plane z = 0 where they meet [Davis, R.O. Elasticity and geomechanics / R.O.
Davis, A.P.S. Selvadurai. Cambridge University press. 1996. 201 p.].
If we consider a spherical surface centered on the origin, we find it is a principal
surface, supporting the minor principal stress σ1 :

3Pz
σ1 = − ,
2πR 3
where R is the sphere radius.
Note how σ1 change the sign from negative to positive values of z, giving tensile
stresses above the median plane z = 0.
Let’s write Kelvin’s solution in the Cartesian coordinate system. First, we consider
the action of the point load applied at the origin and directed parallel to the x 3 axis
(Fig. 8.8).
The point load can be written as below:

X i = Pδi3 · δ(→
x),

where δ(→
x) is the Dirac delta function.
Kelvin’s solution for this case has the form:
( )
(3) λ+μ xi · x3 λ + 3μ δi3
ui = P + · , (8.43)
8π μ(λ + 2μ) R3 λ+μ R

where u i(3) are displacement components caused by the action of a point load directed
along the x 3 axis and applied at the origin R(x, 0) = (xi · xi )1/ 2 .
A solution of formula (8.43) is often called an elementary solution of the first
kind. Formula (8.43) in the literature is often recorded as:
( )
δi3
u i(3) = P A B · − (R),i3 , (8.44)
R

λ+μ
where A = 8π μ(λ+2μ) , B = 2(λ+2μ)
(λ+μ)
.
Stresses and dilatations corresponding to the displacements defined by formula
(8.44) are given by such formulas:
220 8 Fundamental Solutions
[ ( ) ]
∂ R−1 ∂R 2 μ
σ11 = P2μA 3 − ,
∂ x3 ∂ x1 λ+μ
∂ R ∂ R ∂ R−1
σ12 = P6μA ,
∂ x1 ∂ x2 ∂ x3
[ ( ) ]
∂ R−1 ∂R 2 μ
σ13 = P2μA 3 + ,
∂ x1 ∂ x3 λ+μ
[ ( ) ]
∂ R−1 ∂R 2 μ
σ22 = P2μA 3 − , (8.45)
∂ x3 ∂ x2 λ+μ
[ ( ) ]
∂ R−1 ∂R 2 μ
σ23 = P2μA 3 + ,
∂ x2 ∂ x3 λ+μ
[ ( ) ]
∂ R−1 ∂R 2 μ
σ33 = P2μA 3 + ,
∂ x3 ∂ x3 λ+μ
2μ ∂ R−1
θ = PA .
λ + μ ∂ x3

Consider a sphere centered at the origin and calculate the value of the forces on
that sphere pi = σi j · n j . We assume that the normal directs inside the sphere (that
is, n j = −x j /R) and P = 1.
Then in accordance with the formulas (8.45), we have:
( )
6μ A x1 x3 6μ A x2 x3 2μ A 3 · x32 μ
p1 = , p2 = , p3 = + . (8.46)
R4 R4 R2 R2 λ+μ

We investigate whether the ball cut from the body will be in equilibrium. To do
this, check whether the equilibrium conditions are met:
∫ ∫ ∫
p1 d S = 0, p2 d S = 0, p3 d S − 1 = 0. (8.47)
S S S

We substitute (8.46) into expressions (8.47) and integrate along the surface of the
ball r = a. As a result, we get that the equilibrium conditions are met.
Now let’s summarize the results for the general case of the point load in unlimited
space.
( j) −

Denote through displacement Ui (→
x, ξ ) caused by the action of a point load


directed parallel to the x j axis and applied at a point ξ , magnitude is P = 1.
By analogy with the function (8.43) and expression (8.44), you can write:
( ( ) ( )
( j) →)
− xi · x j λ + 3μ δi j B · δi j
Ui → ξ = A
x, + · = A − (R) ,i j .
R3 λ+μ R R
8.6 Cerrutti’s Problem 221

( j) −

Functions Ui (→
x, ξ ) are form the Green’s displacement tensors. Recall that the
Green’s displacement tensor is symmetrical, that is:

→ (−
→ )
( j)
x, ξ ) = U (ij ) ξ , x→ ,
Ui (→



and at a point ξ , its components have a feature of order 1/R, and their values decrease
with increasing values R.


If a mass force X i (ξ) · d V (ξ) is acting at a point ξ , the displacements caused
by that force at the point x→ will be:
[ ]
u i = X 1 (ξ)Ui(1) (x, ξ) + X 2 (ξ)Ui(2) (x, ξ)+ X 3 (ξ)Ui(3) (x, ξ) d V (ξ)
( j)
= X j (ξ)Ui (x, ξ)d V (ξ).

We integrate in the area V ' in which mass forces operate. As a result, we get the
following formula: (−
∫ →) ( j) ( − →)
u i (→
x) = X j ξ Ui x, → ξ d V (ξ) or u i (→
x) =
∫ ( −
→ ) (−→ ) V'
X j ξ U (ij ) ξ , x→ d V (ξ).
V'

8.6 Cerrutti’s3 Problem

Cerrutti’s problem is the problem about a horizontal point load acting on the surface
of an elastic half-space (Fig. 8.10). We consider the problem of loading the surface
x 3 = 0, which bounds the elastic half-space, with tangent forces. Tangent load is
the point load P · δ(x1 )δ(x2 ) and acts at the origin of coordinates, pointing in the x 1
direction.
The point load is represented by P and acts at the origin of coordinates, pointing
in the x 1 direction. This is a more complicated problem than either the Boussinesq’s
or Kelvin’s problem due to the absence of radial symmetry enjoyed by those two
problems. We use a rectangular coordinate system. Cerrutti’s solution is:
[ ( )]
P 1 x12 1 x12
u1 = + 3 + (1 − 2ν) − ,
4π μ R R R + x3 R(R + x3 )2
[ ]
P x1 x2 x1 x2
u2 = − (1 − 2ν) ,
4π μ R 3 R(R + x3 )2
[ ]
P x1 x3 x12
u3 = + (1 − 2ν) . (8.48)
4π μ R 3 R(R + x3 )2

3 Cerrutti V. (1850–1909), Italian scientist-mechanic.


222 8 Fundamental Solutions

Fig. 8.10 Cerrutti’s problem P


x1

x2 x3

[ ( )]
P x1 3x12 (1 − 2ν) 2Rx22
σ11 = − + R − x2 −
2 2
,
2π R 3 R2 (R + x3 )2 (R + x3 )
[ ( )]
P x1 3x22 (1 − 2ν) 2Rx12
σ22 = − + 3R 2
− x 2
− ,
2π R 3 R2 (R + x3 )2
1
(R + x3 )
3P x1 x32
σ33 =− ,
2π R 5
[ ( )]
P x2 3x12 (1 − 2ν) 2Rx12
σ12 = − + −R + x1 +
2 2
,
2π R 3 R2 (R + x3 )2 (R + x3 )
3P x1 x2 x3 3P x12 x3
σ23 =− , σ31 =− , (8.49)
2π R 5 2π R 5
( )1 2
where R = x12 + x22 + x32 / .
Analysis of solutions (8.48–8.49) shows that displacements and stresses at the
origin are singular, while for large R, all components approach to zero. The vertical
plane that contains the x 1 axis is a plane of symmetry. Looking at the x 1 -component of
the displacement field, we see that particles are displaced in the direction of the point
load. The x 2 -component of displacement moves particles away from the x 1 axis for
positive values of x 1 and toward the x 1 axis for negative x 1 . Vertical displacements
take the sign of x 1 and hence particles move downward in front of the load and
upward behind the load. Both the displacements and stresses for Cerrutti’s problem
appear more complex than for either the Boussinesq’s or Kelvin’s problem, but this
is largely an illusion. Cerrutti’s solution is embedded in the Kelvin’s solution for the
special case where ν = 0.5. If we set ν equal to 0.5 in both solutions, we find the
stresses and displacements in the left- or right-hand half-space of Kelvin’s problem
are identical to the Cerrutti’s solution, provided we to designate the coordinate axes
appropriately.
8.7 General Case of Point Load Action on Surface of Falf-Space 223

8.7 General Case of Point Load Action on Surface


of Falf-Space


Consider the general case of loading action: an arbitrary point load P acts at the
origin. Force components are (P1 , P2 , P3 ) and they are directed along three respective
coordinate axes (x 1 , x 2 , x 3 ).
We build a total solution, adding together the Boussinesq’s solution and Cherutti’s
solution. Therefore, the total displacements are obtained by adding the displace-
ments expressed by the formulas (8.7) for the component of the load P3 , and the
displacements caused by the forces P1 and P2 (formulas (8.48)):
[ ( )]
1 x12 1 x12
4 π μ u 1 = P1 + 3 + (1 − 2ν) −
R R R + x3 R(R + x3 )2
[ ] [ ]
x1 x2 x1 x2 x1 x3 x1
+ P2 − (1 − 2ν) + P3 − (1 − 2ν) ,
R3 R(R + x3 )2 R3 R(R + x3 )
[ ]
x1 x2 x1 x2
4π μ u 2 = P1 − (1 − 2ν)
R3 R(R + x3 )2
[ ( )] [ ]
1 x2 1 x22 x2 x3 x2
+ P2 + 23 + (1 − 2ν) − + P 3 − (1 − 2ν) ,
R R R + x3 R(R + x3 )2 R3 R(R + x3 )
[ ]
x1 x3 x12
4 π μ u 3 = P1 + (1 − 2ν)
R3 R(R + x3 )2
[ ] [ ]
x2 x3 x22 x32 1
+ P2 + (1 − 2ν) + P3 + 2(1 − ν) ,
R3 R(R + x3 )2 R3 R

( )1 2
where R = x12 + x22 + x32 / .
If distributed loads p(x1 , x2 ) ≡ ( p1 (x1 , x2 ), p2 (x1 , x2 ), p3 (x1 , x2 )) act on the
plane × 3 = 0 in the area D, then the displacements caused by them can be found using
the superposition principle. Therefore, for example, the displacement component
u 1 (→
x) is expressed by the following formula:

4 π μ u1 (→ x) =
¨ ⎧ [ ( )]
1 (x1 − ξ1 )2 1 x 1 − ξ1
p1 (ξ1 , ξ2 ) + + (1 − 2ν) −
R R3 R + x3 R(R + x3 )2
D
[ ]
(x1 − ξ1 )(x2 − ξ2 ) (x1 − ξ1 )(x2 − ξ2 )
+ p2 (ξ1 , ξ2 ) − (1 − 2ν)
R3 R(R + x3 )2
[ ]⎧
(x1 − ξ1 )x3 (x1 − ξ1 )
+ p3 (ξ1 , ξ2 ) − (1 − 2ν) dξ1 dξ2 .
R3 R(R + x3 )
224 8 Fundamental Solutions

x1

x2 h

x3

Fig. 8.11 Mindlin’s problem

8.8 Mindlin’s4 Problem

Mindlin’s problem is the problem of a point load (arbitrary direction in space) acting
in the interior of an elastic half-space. The load P acts at a point, a distance h beneath
the half-space surface (Fig. 8.11). Plane x3 = 0 is free from external forces.
Mindlin’s problem is a generalization of the Kelvin’s, Bussinesq’s, and Cherutti’s
problems.
As the first problem, consider a special case where the load P acts at the point
(0, 0, h) and directed in the positive direction of the x 3 axis.
We divide the solution of the formulated problem into two stages.
First, we consider the problem of the action in unlimited space of two oppositely
directed forces: the force P = +1 applied at a point (0, 0, h) and the force P = −1
acting at{a point}(0, 0, −h). We denote the stress–strain state corresponding to this
load as u i0 , σi0j . To determine it, we use the Kelvin’s method. We use formulas
(8.44) and apply the principle of superposition. As a result, for displacements u i0 we
get the following formulas:
( ( ) )
xi · (x3 − h) xi (x3 + h) 1 1
u i0 = P A − + B − δi3 ,
R13 R23 R1 R2

where
λ+μ λ + 3μ
A= ,B = ,
8 π μ(λ + 2μ) λ+μ
[ ]1 2 [ ]1 2
R1 = r 2 + (x3 − h)2 / , R2 = r 2 + (x3 + h)2 / ,
r 2 = x12 + x22 .

By defining the stress field σi0j , we can see that in the plane x3 = 0, only the stress
σ33
0
is not zero:

4 Raymond David Mindlin (1906–1987), American mechanical scientist.


8.8 Mindlin’s Problem 225
( )
h 1 − 2ν 3h 2 λ ( )1 2
σ33
0
= + 5 ,ν = , R0 = h 2 + x12 + x22 / .
4π(1 − ν) R03 R0 2(λ + μ)

Therefore, in the second stage, we determine the displacements u i' caused by the
action of a load p3 (x1 , x2 ) = −σ33
0
(x1 , x2 , 0) in the plane x3 = 0 of the elastic
half-space. This load is applied to balance the loads in the plane x3 = 0. This is the
problem discussed in Sects. 8.2 and 8.3 (the Bussinesq’s and Hertzs’ problems).
Finally, formulas for complete displacements have the form:
( ) [ ]
Pr x3 − h x3 + h P ∂. ∂.
ur = − − x3 + (1 − 2ν) ,
16 π μ (1 − ν) R13 R23 4π μ ∂r ∂r
[ ( )]
P (x3 − h)2 (x3 + h)2 1 1
u3 = − + (3 − 4ν) −
16π μ (1 − ν) R13 3
R2 R1 R2
1−ν x3 ∂ .
+ .− ,
2π μ 4π μ ∂ x3

where u r = u 1 cos θ + u 2 sin θ;

1 h x3 + h
.(→
x) = + , x3 > 0.
R2 2(1 − ν) R23
h 1 ∂.
.(→
x) = ln(R2 + x3 + h) − , = ..
2(1 − ν) R2 ∂ x3

The components of stresses in cylindrical coordinate system (r, θ, z) have such


representations:
[ 2
P 3r z 1 (1 − 2ν)z 1 (1 − 2ν)z − 12(1 − ν)h
σrr = − − +
8π(1 − ν) R3 5
R33 R3
]
3r 2 z − 6(7 − 2ν)hz 2 + 24h 2 z 30hz 2 (z − h)
− 5
− + σrr
B
,
R R7
P
σθθ =
8π (1 − ν)
[ ]
(1 − 2ν)z 1 (1 − 2ν)(z + 6h) −6(1 − 2ν)hz 2 − 64h 2 z
− + − + σθθ
B
,
R33 R3 R5
[ 3
P 3z 1 (1 − 2ν)z 1 (1 − 2ν)(z − 2h)
σzz = − + −
8π(1 − ν) R3 5
R33 R3
]
3z 3 + 12(2 − ν)hz 2 − 18h 2 z 30hz 2 (z − h)
− + + σzz
B
,
R5 R7
[ 2
Pr 3z 1 (1 − 2ν) (1 − 2ν)
σr z = σzr = − + −
8π(1 − ν) R3 5
R33 R3
226 8 Fundamental Solutions
]
3z 2 + 6(3 − 2ν)hz − 6h 2 30hz(z − h)
− + + σrBz ,
R5 R7

σr θ = σθr = σθz = σzθ = 0,


[ ]1 2 [ ]1 2
where z 1 = x3 − 2h, z 1 = x3 − 2h, R = r 2 + x32 / , R3 = r 2 + (x3 − 2h)2 / ,
r 2 = x12 + x22 , σ B is the stress component of the Boussinesq’s solution (8.10).
As the next problem, let’s consider the Mindlin’s solution for a horizontal point
load P acting at the point (0, 0, h) and directed in the positive direction of the x 1
axis.
As with the previous problem, in this case, the solution can be represented as
a superposition of two parts, after we use the Kelvin’s solution and Bussinesq’s
solution. But we can use the Cerruti’s solution together with Bussinesq’s solution.
As a result, the additional displacements and stresses are:

P
u x1 =
16 π μ (1 − ν)
( 2 )
x1 3 − 4ν 3 − 4ν −x12 + 2h(x3 − h) 6hx12 (x3 − h)
+ − + −
R33 R3 R R3 R5
( )
P x1 x2 x1 x2 6hx1 x2 (x3 − h)
u x2 = − − ,
16 π μ (1 − ν) R33 R3 R5
( )
P x1 x3 x1 x3 + 2hx1 (3 − 4ν) 6hx1 x3 (x3 − h)
u x3 = − − ;
16 π μ (1 − ν) R33 R3 R5
[ 2
P x1 3x1 (1 − 2ν) (1 − 2ν)
σx1 x1 = − + −
8π(1 − ν) R35 R33 R3
]
3x12 − 6(3 − 2ν)hx3 + 18h 2 30hx12 (x3 − h)
− − ,
R5 R7
[ 2
P x1 3x2 (1 − 2ν) (1 − 2ν)
σx2 x2 = − − +
8π(1 − ν) R3 5
R33 R3
]
3x22 − 6(1 − 2ν)hx3 + 6h 2 30hx22 (x3 − h)
− − ,
R5 R7
[ 2
P x1 3z 1 (1 − 2ν) (1 − 2ν)
σx3 x3 = − + −
8π(1 − ν) R3 5
R33 R3
]
3x32 + 6(1 − 2ν)hx3 + 6h 2 30hx32 (x3 − h)
− − ,
R5 R7
[ 2
P x2 3x1 (1 − 2ν) (1 − 2ν)
σx1 x2 = σx2 x1 = − + −
8π(1 − ν) R3 5
R33 R3
8.9 Some Generalizations 227
]
3x12 − 6h(x3 − h) 30hx12 (x3 − h)
− − ,
R5 R7
[ ]
P x1 x2 3z 1 3x3 + 6(1 − 2ν)h + 18h 2 30hx3 (x3 − h)
σx2 x3 = σx3 x2 = − − − ,
8π(1 − ν) R35 R5 R7
[ 2
P 3x1 z 1 (1 − 2ν)z 1 (1 − 2ν)(x3 − 2h)
σx1 x3 = σx3 x1 = − + −
8π(1 − ν) R35 R33 R3
]
3x1 x3 + 6(1 − 2ν)hx1 − 6hx3 (x3 − h) 30hx1 (x3 − h)
2 2 2
− − .
R5 R7

For the complete solution, we must add the Boussinesq’s solutions (8.6) and (8.10)
to them.

8.9 Some Generalizations

We will make some generalizations regarding the classical fundamental solutions of


the theory of elasticity, which were discussed in this chapter. Classic fundamental
solutions can be grouped into two classes.
The first class of fundamental solutions is the problem for an infinite elastic body.
That is, in this case, the area under consideration .∗ is an infinite elastic medium
and its corresponding boundary . ∗ is an infinite boundary.
In general, this class corresponds to the fundamental Kelvin’s solution for an
infinite elastic medium.
Kelvin’s solution for 3D-problems has the form:

1 [ ]
u i∗j (ξ, x) = (3 − 4ν)δi j + r,i r, j . (8.50)
16π(1 − ν)μ r

Kelvin’s solution for 2D-problems has the form:

−1 [ ]
u i∗j (ξ, x) = (3 − 4ν) ln(r )δi j − r,i r, j , (8.51)
8π(1 − ν)μ

where u i∗j (ξ, x) are displacements occurring at the point x in the jth direction and
corresponding to a point load acting in the ith direction and applied at the point ξ.
Expressions for stresses are written as:
⎧ ⎧
−1 [ ]∂ r ( )
pi∗j (ξ, x) = (1 − 2ν)δi j + β r ,i r , j −(1 − 2ν) r ,i n j − r , j n i ,
4α π (1 − ν)r α ∂n

where α = 2, 1, β = 3, 2 respectively, for 3D-problems and plane-deformed state


problems.
228 8 Fundamental Solutions

Function r = r (ξ, x) is the distance between the point ξ to which the load is
applied and the point x. The derivatives of function r are taken by the coordinates of
the point x, i.e.,
√ / /
r= ri ri , ri = xi (x) − xi (ξ), r,i = ∂r ∂ xi (x) = ri r .

The expressions for deformations ε∗jk at an arbitrary point x due to the point load
applied at the point ξ and directed along the ith axis have the form:

−1 { ( )
ε∗jki (ξ, x) = α
(1 − 2ν) r,k δi j + r, j δik
8α π (1 − ν)μr
( )}
− r,i δ jk + β r,i r, j r,k .

Formulas for stresses can be written in the following form:

−1 { ( ) }
σ∗jki (ξ, x) = α
(1 − 2ν) r,k δi j + r, j δik − r,i δ jk + β r,i r, j r,k .
4α π (1 − ν)r
(8.52)

The expressions for the plane-deformed state are also valid for the plane-stressed
state, if the coefficient ν is replaced by ν = ν/(1 + ν).
The second class of fundamental solutions relates to problems for half-space. In
this case, a semi-infinite medium is considered with a plane section of the boundary—
an infinite horizontal plane .. Typically, a lower half-space is considered.
This class of solutions can be represented in the following form:

()∗ = () K + ()C ,

where () K and ()C are, respectively, Kelvin’s solutions (expressions (8.50−8.52) for
three- and two-dimensional solutions) and an additional solution.
For example, expressions of additional solutions ()C for displacements due to
a point load applied within a half-space, have the form [Brebbia, C.A. Boundary
element techniques/C.A. Brebbia, J.C.F. Telles, L.C. Wrobel. Springer-Verlag Berlin,
Heidelberg, 1984. 524 p.]:
(( )/ ( )/ / )
u c11 = K d 8(1 − ν)2 − (3 − 4ν) R + (3 − 4ν)R12 − 2cx R 3 + 6cx R12 R 5 ,

( / / / )
u c12 = K d r2 (3 − 4ν)r1 R 3 − 4(1 − ν)(1 − 2ν) (R(R + R1 )) + 6cx R1 R 5 ,

( / / / )
u c21 = K d r2 (3 − 4ν)r1 R 3 + 4(1 − ν)(1 − 2ν) (R(R + R1 )) − 6cx R1 R 5 ,

( / / ( / )( / )
u c22 = K d 1 R + (3 − 4ν)r22 R 3 + 2cx R 3 1 − 3r22 R 2
8.9 Some Generalizations 229
( / )( / ))
+ 4(1 − ν)(1 − 2ν) (R + R1 ) 1 − r22 (R(R + R1 )) ,

( / /( ) / )
u c23 = K d r2 r3 (3 − 4ν) R 3 − 4(1 − ν)(1 − 2ν) R(R + R1 )2 − 6cx R 5 ,

( / / ( / )( / )
u c33 = K d 1 R + (3 − 4ν)r32 R 3 + 2cx R 3 1 − 3r32 R 2
( / )( / ))
+ 4(1 − ν)(1 − 2ν) (R + R1 ) 1 − r32 (R(R + R1 ) ,

/ /
u c13 = u c12 r3 r2 ; u c31 = u c21 r3 r2 ; u c32 = u c23 ,

where i = 1, 2, 3; R( =) R(Ri Ri )1/2 , ri = xi (x) − xi (ξ); K d = 1/(16π(1 − ν)μ),


Ri = xi (x) − xi ξ' ; c = x1 (ξ) ≥ 0 , x = x1 (x) ≥ 0.
The geometry of problem about the point load (|P1 | = |P2 | = |P3 | = 1 ) applied
within the half-space is shown in Fig. 8.12.
An additional part of the displacements for the plane-deformed state can be written
in the following form [Brebbia, C.A. Boundary element techniques/C.A. Brebbia,
J.C.F. Telles, L.C. Wrobel. Springer-Verlag Berlin, Heidelberg, 1984. 524 p.]:
{ ( ) ( )/ / }
u c11 = K d − 8(1 − ν)2 − (3 − 4ν) lnR + (3 − 4)R12 − 2cx R 2 + 4cx R12 R 4 ,

с

x2
c
r2 = R2 R1

x3
P2 x
P3
r3 = R3 P1
R

r1
r

x
x1

Fig. 8.12 The geometry of problem about the point load (|P1 | = |P2 | = |P3 | = 1) applied within
the half-space
230 8 Fundamental Solutions

Fig. 8.13 The geometry of ‘


problem about the point load с x2
(|P1 | = |P2 | = 1) applied
within the half-plane
R1 R c

P2
x
P1
r1 r
x
x1
r2 = R2

{ / / }
u c12 = K d (3 − 4ν)r1r2 R 2 + 4cx R1r2 R 4 − 4(1 − ν)(1 − 2ν)θ ,

{ / / }
u c21 = K d (3 − 4ν)r1r2 R 2 + 4cx R1r2 R 4 − 4(1 − ν)(1 − 2ν)θ ,

{ ( ) ( )/ / }
u c22 = K d − 8(1 − ν)2 − (3 − 4ν) lnR + (3 − 4ν)r22 + 2cx R 2 − 4cxr22 R 4 ,
/ /
where i = 1, 2, θ = arctgR2 R1 , K d = 1 (8π(1 − ν)μ), other symbols are
shown in Fig. 8.13.
Expressions for additional stresses due to the point loads applied inside the half-
plane for the plane-deformed state are given. /
For a plane-stressed state, the coefficient ν is replaced by ν = ν (1 + ν) in the
formulas.
It follows from expressions for additional terms that they do not contain singular-
ities at x1 ≥ 0 and c > 0 (that is, when the point of loading application is inside the
area .∗ ). When the point of application of the load lies on the surface .(c → 0),
then additional expressions, together with the Kelvin’s solution, give a complete
solution to the three-dimensional Bussinesque-Cherruti problem or two-dimensional
Flaman’s problem.
In the case of the Flaman’s problem,
( for )example, the fundamental displacements
and stresses on the surface equal to ξ ∈ . :
{ } ∗ { }
u ∗11 = K d' 2(1 − ν)lnr − r,12
, u 12 = −K d' (1 − 2ν)θ − r,1r,2
{ } { }
u ∗21 = −K d' −(1 − 2ν)θ − r,2 r,1 , u ∗22 = −K d' 2(1 − ν)lnr − r,2 2
,
( )
2 ∂r
pi∗j = − r,i r, j (8.53)
πr ∂n
/
where K d' = 1 (2π μ).
8.9 Some Generalizations 231

Once again, some comments on such an important question as the transition from
a three-dimensional problem to a plane-deformed state problem. It should be noted
that in Sect. 8.4 this problem was already discussed regarding the Flaman’s solution.
In three-dimensional displacement problems, the Kelvin’s solution u i∗j tends to
zero at r → ∞. In two-dimensional problems, this is not the case, since u i∗j → −∞
at r → ∞ due to the presence of a logarithmic function in the expression for a
fundamental solution with a flat-deformed state.
This behavior in the two-dimensional case does not represent anything unex-
pected. Therefore, from a physical point of view, one can consider the case of a
semi-infinite rod, the x coordinate of which varies from x(A) = 0 to x(B) → ∞.
We assume that the end B is rigidly fixed, and at the point A we apply an axial load
in the positive direction. As a result, constant deformations are obtained in the rod.
Then, after integration, we get that u( A) → ∞. On the other hand, if the displace-
ment at the point A is counted from a point C at a finite distance from the point A,
then the displacement u(A) will be finite and u(C) = 0, u(B) → −∞.
This simple example clearly shows the physical nature of the expression (8.51) and
can be used to explain the transition from a three-dimensional to a two-dimensional
problem by integration of the coordinate x3 (ξ).
Algorithms of integration of fundamental solutions under action of loads on arbi-
trary areas of influence are described, for example, in [Davis, R.O. Elasticity and
geomechanics / R.O. Davis, A.P.S. Selvadurai.—Cambridge University press. 1996.
201 p.].

Control Questions

1. The fundamental solution of the Laplace equation for 3D-problem:


(1) K (x, y) = − 4πr
1
;
(2) K (x, y) = 2π ln r1 ;
1
( )
(3) K (x, y) = exp r1 .
2. What is the mathematical statement of the Boussinesq’s problem?
3. Where the Bussinesq’s solution has a singular feature?
4. What is the Hertz’s task?
5. What is the mathematical statement of the Flamant’s problem for a half-space?
6. What are the difficulties and features of building up the Flamant’s solution for
a half-plane?
7. What is the mathematical statement of the Kelvin’s problem?
8. What is the definition of the Green’s displacement tensor?
9. What is the mathematical statement of the Cerrutti’s problem?
10. What is the difference in the mathematical statement of the Mindlin’s problem
and Kelvin’s problem?
Chapter 9
Dynamic Problems of Solid Mechanics

9.1 General Concepts and Definitions

In previous chapters, we considered the solid mechanics problems in a static formula.


However, there is a large class of problems where the external loads on deformable
bodies are clearly dynamic. In this chapter, we consider problems whose model
analogues related to wave problems of the mechanics of deformable solids.
In short, we give some basic concepts related to the description of wave processes
in solids.
If oscillations of particles are induced in some places of the elastic solid medium,
then, due to the interaction among the particles, these oscillations propagate from the
particle to the particle at the velocity v. The process of propagating fluctuations in
space is called a wave. It is important that such disturbances, which occur in some
parts of the elastic solid medium, are propagating in it at a finite velocity. This means
that a point at a distance L from the source of disturbances, for the time τ, a rest
will be maintained. Thus, the particles of the medium in which the wave propagates
are not involved by the wave in the translational movements, and they only oscillate
near their equilibrium positions. The characteristic property of the wave is the
transfer of the energy without the transfer of the matter.
The concepts of “oscillations” and “waves” are interconnected, but also different
from each other. The motion of the body G is called the oscillatory motion (or the
simply oscillations) at the finite time interval [t1 , t2 ] (at the semi-infinite interval t ≥
t1 ), if at this interval all points M ∈ G oscillate. The oscillations of the point M in the
interval [t1 , t2 ] is meant such a law of change in time of at least one defining parameter
u i (M, t) for which there exists a function u i∗ (M, t) with a trajectory more than two
(counting set) points of intersection, that is, the equation u i (M, t) = u i∗ (M, t) has
more than two (counting set) roots (Fig. 9.1). Usually, u i∗ (M, t) corresponds to the
position of equilibrium of the body.
From the point of view of such definition of the oscillatory motion, if there are
points in the body G that oscillate with different laws of motion, then such a movement
is called a wave process (waves).

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 233
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_9
234 9 Dynamic Problems of Solid Mechanics

Fig. 9.1 Point oscillation (counting set of roots with u i∗ )

Once again, it should be emphasized that the fundamental difference between


waves and oscillations is the energy transfer, since only “local” energy transforma-
tions occur in oscillations. Waves, as a rule, are able to move long distances from
their places to the origin.

Remark Most waves are by nature not new physical phenomena, but only a condi-
tional name for a certain type of collective movement of medium particles. Most
waves are oscillations of some media. Outside this medium, waves of this type do
not exist (for example, sound in a vacuum).

Wavelength is the distance over which a wave propagates in one period of


oscillation.
Depending on the direction of oscillation of the particles with respect to the
direction in which the wave propagates, longitudinal and transverse waves are
distinguished.
Longitudinal waves are waves in which the oscillations of the medium are parallel
(“along”) to the direction of the wave traveling and the displacement of the medium is
in the same (or opposite) direction of the wave propagation. Mechanical longitudinal
waves are also called compressional or compression waves, because they produce the
compression and rarefaction when traveling through a medium, and pressure waves,
because they produce increases and decreases in pressure.
The other main type of wave is the transverse wave, in which the displacements of
the medium are at right angles to the direction of propagation. Transverse waves, for
instance, describe some bulk sound waves in solid materials (but not in fluids); these
are also called “shear waves” to differentiate them from the pressure (longitudinal)
waves that these materials also support.
9.1 General Concepts and Definitions 235

If the relationship between the particles of the medium is made by the elastic
forces, resulting from the deformation of the medium during the transfer of
oscillations from one particle to another, then the waves are called elastic waves.
Elastic transverse waves can only occur in a medium having shear resistance.
Therefore, only longitudinal waves are possible in liquid and gaseous media. Both
longitudinal and transverse waves can occur in the solid medium.
In fact, not only particles along the x axis oscillate in solids, but also a collection
of particles enclosed in a certain volume. Propagating from the source of oscillations,
the wave process covers more and more parts of solid. The geometric location of
the points to which the oscillations reach the time t is called the wave front. The
wave front is the surface that separates the part of space already involved in the wave
process from the area in which the oscillations have not yet occurred. The geometric
location of points oscillating in the same phase is called the wave surface. The wave
surface can be passed through any point of the space covered by the wave process.
Therefore, there is an infinite set of wave surfaces, while the wave front at each point
in time is only one. Wave surfaces remain stationary. Wave front displaces all the
time. Wave surfaces can be any shape. If the wave surfaces are planes, then the waves
are called plane waves. Spherical waves are waves whose wave surfaces are spheres.
Waves that have breaks derived from displacements (i.e., velocity and stress
breaks) are called strong breaking waves or shock waves. Shock waves occur during
explosions, detonation, supersonic movements etc.
Two and only two types of elastic waves can propagate in an unlimited isotropic
solid: longitudinal and transverse waves.
In addition to the above types of waves that can propagate within the limitless
solids, elastic waves can propagate along the surface as well. At the same time,
the displacements associated with these waves decrease with the depth according
to the exponential law. These waves are called Rayleigh surface waves and play an
important role, for example, in the study of seismic phenomena. Since these waves
diverge in only two dimensions, they attenuate with a distance slower than other
types of waves. The waves on the surface of the body are neither longitudinal nor
transverse.
Surface waves in solids are of two classes: with vertical polarization, in which
the vector of oscillatory displacement of medium particles is located in a plane
perpendicular to the boundary surface and with horizontal polarization, in which the
vector of displacement of medium particles is parallel to the boundary surface.
The most common particular cases of surface waves include the following.
1. Rayleigh waves propagating along the boundary of a solid with a vacuum or a
sufficiently rarefied gas medium. The energy of these waves is localized in the
surface layer with a thickness of l to 2l, where l is the wavelength.
2. Decaying waves of the Rayleigh type at the boundary of a solid with a liquid,
provided that the phase velocity in the liquid is less than in the solid (which is
true for almost all real media). This wave continuously emits energy into the
liquid, forming in it an inhomogeneous wave extending from the boundary. The
236 9 Dynamic Problems of Solid Mechanics

depth distribution of displacements and stresses is the same as in the Rayleigh


wave.
3. Stoney wave1 propagating along the flat boundary of two solid media, in the case
where the modulus of elasticity and density of which do not differ much. Such a
wave can be represented as the sum of two Rayleigh waves—one in each media.
4. Love waves are surface waves with horizontal polarization that propagate at the
boundary of the solid half-space with the solid layer. These waves are purely
transverse: they have only one displacement component, and elastic deformation
in the Love wave is a pure shift.
An important property of waves is the superposition property, according to which
in any region of the medium any number of waves can propagate independently from
each other.
In the theory of elasticity, the body is considered to be in equilibrium under the
influence of applied forces. It is assumed that elastic deformations have already taken
their static values. Such an interpretation is exact enough for problems in which the
period between the moment of loading application and the establishment of actual
equilibrium is small compared to the time periods during which observations are
made. However, when the effect of forces applied only for a short period of time or
rapidly changing is examined, this phenomenon must be considered from the point
of view of the propagation of stress waves.
Real bodies are never completely elastic. Therefore, when disturbances propagate
in solids, part of the mechanical energy turns into heat (several different mechanisms
of these transformations are combined by the common name, namely, internal fric-
tion). The velocity gradients produced by the stress wave cause a second type of
loss due to the viscosity of the material. The nature of the attenuation for these two
types of internal friction is different. Experimental data show that both types occur
in solids.

9.2 Wave Equation

The one-dimensional wave equation can be represented in the following form:


• In Cartesian coordinates:

∂ 2ϕ ∂ 2ϕ
= a2 2 ; (9.1)
∂x 2 ∂t

• In spherical coordinates (r is radial coordinate):

∂ 2 ψ 2 ∂ψ ∂ 2ψ
+ = a2 2 . (9.2)
∂r 2 r ∂r ∂t

1 George Johnstone Stoney (1826–1911), British, Irish physicist and mathematician.


9.3 Model Problems of Dynamics of Elastic Body 237

Remark Equations of this type are considered in the course of the equations of
mathematical physics. The main feature of the solutions of the wave equation is the
description of a perturbation propagating with a finite velocity, i.e., a wave.

Using the Dalamber method, one can consider the process of propagation of waves
(perturbations) of the type:

ϕ(x, t) = Φ 1 (x − t/a) + Φ 2 (x + t/a), (9.3)

where Φ 1 , Φ 2 are twice continuously differentiable functions of their arguments.


The expression (9.3) represents a general solution to Eq. (9.1). It can be shown
that Eq. (9.1) does not have solutions other than those that can be represented by
functions of the form (9.3) or by a superposition of such functions.
If ax − t = const, , then there is a wave (disturbance) propagating towards the
increase of the ordinate x (when increasing t). If ax + t = const, then there is a wave,
spreading in the opposite direction.
Using (9.3), the properties of the spherical symmetric wave described using the
Eq. (9.2) can be investigated:

r ψ(r, t) = Ψ1 (r − t/a) + Ψ2 (r + t/a). (9.4)

In (9.4), the first term Ψ1 (r − t/a) describes a diverging wave, and the second term
Ψ2 (r + t/a) describes a converging wave. Therefore, if a certain value of the function
ϕ(x, 0) is specified at the beginning of time (t = 0), then it causes perturbations that
can be detected when t > 0 in the vicinity of points x1 and x2 .
That is, the feature of Eq. (9.1) is that it is possible to describe perturbations
(waves) propagating at a finite velocity by this equation.

9.3 Model Problems of Dynamics of Elastic Body

9.3.1 General Comments

Consider various representations of the equations of dynamics of homogeneous


isotropic elastic body.
The equations of motion of an elastic body can be obtained from equilibrium equa-
tions by adding inertia forces to the acting mass forces, according to the D’Alembert
principle ρF i , i.e., ρFi → ρ(Fi − ü i ).
Thus, the equations of motion can be represented in the following form:

σi j, j + ρ(Fi − ü i ) = 0, (i, j = 1, 2, 3). (9.5)

In the expanded form, the Eq. (9.5) is written as:


238 9 Dynamic Problems of Solid Mechanics

∂σx ∂τx y ∂τx z ∂ 2u


+ + +X =ρ 2,
∂x ∂y ∂z ∂t
∂τx y ∂σ y ∂τ yz ∂ 2v
+ + +Y =ρ 2, (9.6)
∂x ∂y ∂z ∂t
∂τx z ∂τ yz ∂σz ∂ 2w
+ + +Z =ρ 2 .
∂x ∂y ∂z ∂t

where X, Y, Z are components of mass forces in the Cartesian coordinate system x,


y, z, and t is time.
The remaining equations of the complete resolution system of equations (equation
of the compatibility of deformations, Cauchy relations, boundary conditions) are
preserved in the same form.
In addition, initial conditions must be added to these equations:

u i = u 0i , u̇ i = u̇ 0i , t = 0. (9.7)

Remark We will not consider the question of proofing the existence and uniqueness
of solving dynamic problems of the theory of elasticity. This question is considered
in detail in the literature on the general course of the dynamic theory of elasticity.
The relationship of stresses with strains when considering an elastic isotropic
body for dynamic problems of the theory of elasticity is determined by the Hooke’s
law:
∂u ∂v ∂w
σx = λε + 2μ , σ y = λε + 2μ , σz = λε + 2μ ,
∂x ∂y ∂z
( ) ( ) ( )
∂u ∂v ∂w ∂u ∂w ∂v
τx y = μ + , τx z = μ + , τx y = μ + ,
∂y ∂x ∂x ∂z ∂y ∂z
∂u ∂v ∂w
ε= + + , (9.8)
∂x ∂y ∂z

where λ, μ are elastic Lamé constants.


As in the construction of the Lamé equations for static problems of the theory
of elasticity, from the system (9.6), we exclude the components σi j of the stress
tensor using the relations (9.8). As a result, we obtain a system of three second-order
equations from displacements u, v, w by coordinates, x, y, z and time t:

∂ε ∂ 2u
(λ + μ) + μΔu + X − ρ 2 = 0,
∂x ∂t
∂ε ∂ 2v
(λ + μ) + μΔv + Y − ρ 2 = 0, (9.9)
∂y ∂t
∂ε ∂ 2w
(λ + μ) + μΔw + Z − ρ 2 = 0.
∂z ∂t
9.3 Model Problems of Dynamics of Elastic Body 239

Equation (9.9) can be written in compact form as follows:

ρü i = (λ + μ)θ,i + μΔu i + ρFi . (9.9*)

The system of Eq. (9.9) is called the Lamé system of dynamic problems of the
theory of elasticity.
In most cases, when considering the dynamic problems of the theory of elasticity,
the Eq. (9.9) are used as the basic ones. Therefore, approaches and methods for
constructing general solutions for dynamic problems of elasticity theory are mostly
developed for a system of resolving equations with basic Eq. (9.9).
It should be noted that in addition to this statement, depending on the boundary and
initial conditions, equations of motion (9.6) relative to stresses are often adopted as
the basic equations of the resolving system of equations. In this case, the system (9.6)
is supplemented in the general case by six Beltrami-Michell equations (compatibility
conditions for stresses).
The Lamé system (9.9) can be presented in several different forms. When
constructing a solution to a system (9.9), it is often convenient, on the basis of
the known theorem of vector analysis (Helmholtz theorem), to represent a vector of
displacements u→ = u = (u, v, w) in the form of potential and solenoidal parts:

→ or u = gradϕ + rotψ,
u→ = gradϕ + rotψ (9.10)
{ }
→ ψx , ψ y , ψz is the vector potential of
where ϕ(x, y, z, t) is the scalar potential, ψ
the displacement field. The “grad” and “rot” operators in a rectangular Cartesian
coordinate system have the form:
| |
|e e e |
∂ϕ | 1 2 3 |
| |
gradϕ = ei , rotψ = | ∂∂x1 ∂∂x2 ∂∂x3 |.
∂ xi | |
| ψ1 ψ2 ψ3 |

Note that for uniqueness of motion potentials, an additional condition is necessary,


which, as a rule, is accepted in the following form:

divψ = ψi,i = 0. (9.11)

The mass force field is also represented as a potential and solenoid component:

F = gradΦ + rotΨ.

.
Note that when using the Eq. (9.10), four functions ϕ, ψx , ψ y , ψz are intro-
duced instead of the three functions u, v, w. It follows, in particular, that there is a
“arbitrariness” in choosing such representation of displacements through potentials.
240 9 Dynamic Problems of Solid Mechanics

Using the representation (9.10), the Lamé equations of motion (9.9) are satisfied
if the displacements potentials are solutions to the following equations (note, that
div grad = δ, div rot = 0):

ϕ̈ = c12 δϕ + Φ , ψ̈ = c22 δψ + ψ, δψ = δψi ei . (9.12)

In (9.12), values
∫ ∫
c1 = (λ + 2μ)/ρ, c2 = μ/ρ, c1 > c2

have a velocity dimension and are accordingly called the velocity of longitudinal
waves (tension–compression waves) and transverse waves (forming waves).
Thus, for a homogeneous elastic isotropic medium, a closed system of motion
equations in displacement potentials consist of Eqs. (9.12) and (9.11). In this case,
the initial conditions (9.7) for it must be recorded in potentials:
|
ϕ|t=0 = ϕ0 , ϕ̇|t=0 = ϕ̇0 , ψ|t=0 = ψ0 , ψ̇|t=0 = ψ̇0 ,

where u 0i ei = gradϕ0 + rotψ0 , u̇ 0i ei = gradϕ̇0 + rotψ̇0 , divψ0 = 0, divψ̇0 = 0.


Note that, despite the independence of Eq. (9.12), as a rule, in initial-boundary
problems they are connected by boundary conditions.
In the absence of mass forces from (9.9), it follows four wave equations for ϕ
→ which can be written as:
and ψ,

2→
∂ 2ϕ → = b2 ∂ ψ ,
Δϕ = a 2 , Δψ (9.13)
∂t 2 ∂t 2

where Δ = ∂∂x 2 + ∂∂y 2 + ∂∂z 2 is the Laplacian, a and b represent, respectively, inverse
2 2 2


√of velocities of longitudinal and transverse waves: a = ρ/(λ + 2μ) = 1 c1 ,
values
b = ρ/μ = 1/c2 .
A similar view is true in the presence of mass forces. In this case, it is necessary
to also represent the mass force vector as a sum of scalar and vector potentials.

9.3.2 Cauchy Problem and Boundary Problems

The fifth chapter has considered the boundary problems of the theory of elasticity and
defined the main types of boundary conditions for the boundary value problems of the
static and dynamic theory of elasticity. In the case of dynamic processes, boundary
conditions must be supplemented by initial conditions that can be formulated for the
elastic displacement vector:
9.3 Model Problems of Dynamics of Elastic Body 241
|
∂u ||
u→|t=t0 = u→0 (x, y, z), = u→1 (x, y, z). (9.14)
∂t |t=t0

Consider the applied and scientifically important Cauchy problem for a system of
Eq. (9.9). The Cauchy problem for the system (9.9) is a problem with initial data
for a limitless elastic medium. Initial conditions (9.14) are set in elastic medium or
in some parts of it. It is necessary to build and investigate the solution of system (9.9)
at t > t0 . From a physical point of view, this is a problem to spread perturbation in
an elastic medium with velocities c1 and c2 . These perturbations are “induced” by
the part of the elastic medium in which at t = t0 functions u→0 (x, y, z), u→1 (x, y, z)
were determined. In the other parts of the medium these functions are zeros.

Remark We will return to the solution of the formulated Cauchy problem further.

Now we turn to the wording of the main boundary value problem in terms of elastic

displacements and in terms of scalar ϕ(x, y, z, t) and vector ψ(x, y, z, t) potentials.
The boundary value problem in the case of boundary conditions in stresses is
formulated as follows (first boundary value problem).
Let the elastic medium occupy some parts of the space D which bounded by the
smooth surface S on which the stresses are set. Mass forces X, Y, Z are defined in
the region D. It is necessary to find a solution to the system of Eq. (9.9), satisfying
the initial conditions (9.14) and boundary conditions on the surface S. The solution
must have a specified smoothness (which is determined by the required smoothness
of the boundary and initial conditions).
The components of the stress vector on the surface, which are included in the
formula of boundary conditions, are expressed by Cauchy formulas through the
components of the stress tensor and the direction cosines of the outer normal line
n→(x, y, z) to the surface S:

Tnx = σx cos(n, x) + τx y cos(n, y) + τx z cos(n, z),


Tny = τx y cos(n, x) + σ y cos(n, y) + τ yz cos(n, z), (9.15)
Tnz = τx z cos(n, x) + τ yz cos(n, z) + σz cos(n, z).

Note that the normal to the surface S is known because the surface S is set and
boundary conditions at moments t ≥ t0 relate to the position of the surface S before
deformations. Therefore, for example, in the case of an elastic half-space z ≥ 0 we
have:

cos(n, x) = cos(n, y) = 0, cos(n, z) = −1.

Then from the conditions (9.15), it is as follows:

Tnx = −τx z , Tny = −τ yz , Tnz = −σz . (9.16)


242 9 Dynamic Problems of Solid Mechanics

That is, on the surface of the half-space in this case, a normal force Tnz = Tzz and
two conditions with respect to tangent forces Tnx , Tny must be set.
The construction of a complete system of resolving equations of the boundary
value problem of the dynamics of an elastic body in the case of boundary conditions
in stresses can be carried out in two ways. The first approach: in boundary conditions
(9.15) we turn to displacements, replacing stresses with their expressions (9.8) in
accordance with the Hooke’s law.
In this case, the boundary conditions of type (9.16) for the half-space, in
accordance with this solution construction, are written as follows:
( ) ( )
∂u ∂w ∂v ∂w
−Tnx = μ + , −Tny = μ + ,
∂z ∂x ∂z ∂y
( )
∂u ∂v ∂w ∂w
−Tzz = λ + + + 2μ . (9.17)
∂x ∂y ∂z ∂z

As a result, the boundary value problem is formulated entirely in terms of displace-


ments: find a solution to the system (9.9) under initial conditions (9.14) and boundary
conditions (9.17), assuming some smoothness of functions u→0 (x, y, z), u→1 (x, y, z),
Tzx (x, y, 0, t),Tzy (x, y, 0, t) and Tzz (x, y, 0, t).
The second approach is to supplement the dynamics Eq. (9.6) with Beltrami-
Michell ratios and formulate the initial conditions (9.14) in terms of stresses. Thus,
in the second case, the problem is fully formulated in terms of stresses.
Finally, we note another possible formula of boundary problems related to the
transition from displacements (or stresses) to scalar and vector potentials, that is,
to the transition from the system (9.9) to the system (9.13). Obviously, in order to
construct a problem, as a basic system (9.13), it is necessary to write both boundary
and initial conditions in terms of potentials.
Initial conditions (9.14) in terms of potentials take the form:
|
ϕ|t=t0 = ϕ0 (x, y, z), ψ→ || → 1 (x, y, z),

t=t0
| |
∂ϕ || ∂ψ→ || (9.18)
= ϕ (x, y, z), | =ψ → 1 (x, y, z),
∂t |t=t0
1
∂t |
t=t0

where ϕ0 , ϕ1 and ψ →0 , ψ
→ 1 are associated with u→0 and u→1 according to ratios (9.10),
which are right with all t ≥ t0 .
When formulating the boundary conditions of the first boundary value problem
in terms of potentials, it is necessary in (9.15) to replace stresses through strains
according to (9.8), and then switch from displacements to derivatives from potentials
according to (9.10). In expanded form, formulas (9.10) have the form:

∂ϕ ∂ψz ∂ψ y ∂ϕ ∂ψx ∂ψz ∂ϕ ∂ψ y ∂ψx


u= + − ,v = + − ,w = + − ,
∂x ∂y ∂z ∂y ∂z ∂x ∂z ∂x ∂y
9.4 Stokes Problem 243

where ψx , ψ y , ψz are components of vector potentials.


In accordance with the remark to formulas (9.10), in the general case, one of the
components of the vector potential or some combinations of functions ϕ, ψx , ψ y ,
ψz can be set to zero. Usually, this fact is used in solving specific problems in which
this arbitrariness is used for the wording (for example, symmetry considerations)
and reduction of calculations.
As an example, the boundary conditions of the first boundary value problem for
the elastic half-space (9.17) are taken (bearing in mind this arbitrariness)ψz = 0:

1 ∂ 2ϕ ∂ 2ψy ∂ 2ψy ∂ 2 ψx
− Tzx = 2 − + − ,
μ ∂ x∂z ∂z 2 ∂x2 ∂ y∂z
1 ∂ 2ϕ ∂ 2 ψx ∂ 2 ψx ∂ 2ψy
− Tzy = 2 + − + ,
μ ∂ y∂ z ∂z 2 ∂ y2 ∂ x∂ y
[ 2 ] (9.19)
1 1 − 2v ∂ ϕ ∂ 2 ψ y ∂ 2 ψx
− Tzz = + − +
μ 2(1 − v) ∂z 2 ∂ x∂z ∂ y∂z
[ 2 ]
2v ∂ ϕ ∂ 2 ϕ ∂ 2 ψx ∂ 2ψy
+ + − .
1 − 2v ∂ x 2 ∂ y2 ∂ y∂z ∂z∂ x

Thus, in terms of potentials, the first boundary value problem is formulated as


follows: it is necessary to construct a solution of the system of Eq. (9.13) under initial
conditions (9.18) and boundary conditions (9.19).

Remark The boundary conditions (9.19) are complex. They are combinations of
second derivatives from potentials through three functions Tzx , Tzy and Tzz known
on the surface of the elastic medium.

The formulas of boundary value problems in the case of displacements set at the
body boundary (the second boundary value problem) and the mixed problem are
performed by analogy with the first boundary value problem.

9.4 Stokes Problem

Dynamic Problems on the Action of Point Forces in Limitless Elastic Medium


The problem of the action of a point force in a limitless elastic medium for problems
of the static theory of elasticity was considered in the corresponding section on
fundamental solutions of the theory of elasticity. Consider a similar problem, but in
the case of the action of a point force, the modulus of which is a function of time.
As with a static problem, the solution for a dynamic problem has a wide range of
applications.
Therefore, let’s at some points in the elastic medium, taken as the origin of coordi-
nates, act a point force P(t), varying in time. For an infinite elastic medium, boundary
conditions at infinity are conditions of no stresses and displacements. Therefore, the
244 9 Dynamic Problems of Solid Mechanics

solution must satisfy the conditions at infinity and have a feature at the origin of
coordinates that leads to a finite point force in magnitude.
As in the case of a static problem, if we cut a small neighborhood in the form of
a ball around the point of application of forces, then stresses are equivalent to the
applied force in magnitude and direction should be distributed over the surface of this
ball. At the same time, it is easy to find (as in the static problem) that at the origin of
the coordinates the stresses should turn to infinity of order r −2 . Then displacements
(since they are expressed through derivatives of stresses and deformations) must have
a peculiarity of order r −1 .
In order to find a class of solutions with a feature of this type, consider the solutions
of the wave equation:

1 ∂2 f
Δf = . (9.20)
c2 ∂t 2

We investigate the solution (9.20) in the case of central symmetry, i.e., f = f (r, t),
r is the radial coordinate, t is time. In this case, the Dalamber operator is simplified
and Eq. (9.20) takes the form:
( )
∂2 2 ∂ 1 ∂2 f 2 ∂2 1 ∂2 f
+ f = or (r f ) = .
∂r 2 r ∂r c2 ∂t 2 r ∂r 2 c2 ∂t 2

Remark The Dalamber operator ◻ a is defined as follows:


( ) ( ) ( )
∂ϕ ∂ϕ 1 ∂ 2 ∂ϕ
◻ a =δ − 2 2 .
∂x ∂x a ∂t ∂ x

From this equation it follows (using the Dalamber integral) the representation of
the solution in the spherical symmetric case:

1 1
f = f 1 (t − r/c) + f 2 (t + r/c),
r r
where f 1 and f 2 are arbitrary functions of specified arguments. The terms f 1 and f 2
describe divergent and convergent spherical waves.
If the action of the force P(t) began at a point in time t = 0, then at t < 0, all the
elastic media were at rest, so that there are no disturbances in the form of converging
waves at t ≥ 0. Therefore, only divergent waves can be considered as the required
solution with the desired feature:
1
f = f 1 (t − r/c). (9.21)
r
Indeed, the function f has the feature at the origin of coordinates exactly those
displacements should have in the point force problem. Note that the derivatives of
function f by coordinates are solutions to the wave equation, but already have terms
9.4 Stokes Problem 245

with a higher feature at the origin of coordinates. Therefore, the first derivatives have
a feature of order r −2 . For example:

∂ ( )
(1/r ) = − 1/r 2 (x/r ),
∂x

where x/r = cos(r, x).


The solution to the Stokes problem has the form [Shemyakin, E.I. Dynamic prob-
lems of the theory of elasticity and plasticity / E.I. Shemyakin. – Moscow. Pub. by
“NNCGP-IGD A.A. Skochinskiy”, 2007. 207 p. (in Russian)]:

( ) ∫τ/ b ⎧ ⎧
∂2 1 x2 1 ( r) 1 ( r)
4πρu = 2 τP(t − τ)dτ + 3 2 P t − − 2P t− ,
∂x r r a a b b
τ/ a

( ) ∫τ/ b ⎧ ⎧
∂2 1 xy 1 ( r) 1 ( r)
4πρv = τP(t − τ)dτ + 3 P t − − P t − ,
∂ x∂ y r r a2 a b2 b
τ/ a

( ) ∫τ/ b ⎧ ⎧
∂2 1 xz 1 ( r) 1 ( r)
4πρw = τP(t − τ)dτ + 3 P t− − 2P t−
∂ x∂z r r a2 a b b
τ/ a

where (u, v, w) are components of the displacement vector. Force P(t) acts in the
axis x direction. Constants “a” and “b” are wave propagation velocities.
Using Stokes formulas, you can reduce the general case of a problem with mass
forces to the case when they are absent.
To do this, we use the linearity of the equation and the boundary conditions of the
problem and sum up two solutions: the general solution of homogeneous dynamic
equations and the partial solution of a heterogeneous system of equations, constructed
using Stokes formulas.
Let mass forces (X, Y, Z ) depending on coordinates and time t act on the inves-
tigated elastic body in a volume V at points (ξ, η, ζ). Denote displacements caused
in a limitless elastic medium at a point (x, y, z) by a force X (ξ, η, ζ, t) acting at a
point (ξ, η, ζ) (the center of gravity of a small volume dω) in the x axis direction
as:

U (x) (ξ, η, ζ, x, y, z), V (x) , W (x) .

Then the force X (ξ, η, ζ, t) causes a displacement equal to:


( (x) )
U dω,V (x) dω, W (x) dω .

And finally summing up the action of all mass forces, we find:


246 9 Dynamic Problems of Solid Mechanics
˚
{ (x) }
U∗ = U + U (y) + U (z) dω, V ∗ = · · · , W ∗ = · · · .
V

These integrals represent elastic displacements in a limitless body due to the action
of mass forces. If the body is bounded, then you need to construct total displacements
in the following form:

u = U ∗ + u ' , v = V ∗ + v ' , w = W ∗ + w'

where the components of displacements u ' , v ' , w ' satisfy homogeneous equations
and so that the boundary conditions on the surface and the initial conditions for
u, v, w are satisfied accurately. Therefore, with given mass forces (and there-
fore with known U ∗ , V ∗ , W ∗ ), the problem of constructing auxiliary functions
u ' , v ' , w ' is a well-defined boundary value problem of the dynamics of an elastic
body.
Thus, one of the important applications of the Stokes solution is the construction
of general solutions to the dynamic equations of the theory of elasticity.

9.5 Natural and Forced Harmonic Oscillations

Consider problems about free (natural) and forced harmonic oscillations, as well as
problems about the propagation of progressive waves in an elastic body.
Natural (free) oscillations. The free oscillation problem refers to the eigenvalue
problem (natural frequency) for homogeneous equations of elasticity theory under
homogeneous boundary conditions. In this case, the elastic body is free from the
action of external forces, that is:

Fi = 0, σi j l j = 0, on Sσ .

Part of the surface S u can be fixed, that is, on it ui = 0. Initial conditions are given
(9.7), as a result of which the body comes into motion.
Forced harmonic oscillations. In this case, the body forces F i , the surface forces
Ri , and the predetermined surface point displacements u0i are periodic time functions
such that:

Fi = Fi0 ϕ(t), Ri = Ri0 ϕ(t), u 0i = u i0 ϕ(t).

Multipliers with the upper index “0” are time-independent, so as a typical


representative of the function ϕ(t), you can take:

ϕ(t) = exp(i pt) = cos pt + i sin pt.


9.5 Natural and Forced Harmonic Oscillations 247

Such representation is true, since any periodic function is represented by a


Fourier series. By building a solution for one member of this series, you can use
the superposition principle to build a complete solution.

9.5.1 Natural Oscillations of Elastic Bodies

Consider the problem of natural oscillations of elastic bodies. We build the solution
of the equations of motion in the form:

u k = Uk exp(i ωt), (9.22)

where U k is the function of coordinate only, not time. In the same way, all components
of the strain tensor εij and stress tensor σij are presented.
In the future, for convenience, through ui , σij , we will denote the amplitudes
of displacements and stresses. That is, instead of U i , we will further write ui . Then,
substituting the required solution (9.22) into the equations of motion (9.6), we obtain
such a system of equations for amplitudes:

σi j , j −ρω2 u i = 0. (9.23)

The external forces and boundary conditions in the case of free oscillations must
be zero, while the factor exp(iωt) is reduced.

Remark We will consider the question of initial conditions later.

Equations that establish a relationship between the amplitudes of stresses and


strains retain the form of ordinary Hooke’s law equation: σi j = E i jkl u k ,l .
The system of Eq. (9.23) under uniform boundary conditions can have an obvious
trivial solution: u i ≡ 0, σi j ≡ 0. However, with some values of the parameter
ω = ωk , a non-zero solution is also maybe: u i ≡ u ik , σi j ≡ σikj .
The corresponding values of the parameter ωk are called the fundamental
frequencies of the elastic body and the functions u ik determine the fundamental
forms of oscillation.
Note that (9.23) includes squares of fundamental frequency, so the root ωk will
always correspond to the second, equal in magnitude and opposite in sign, root (−ωk ).
We will not introduce special numbering for these negative roots, but it should be
remembered that in addition to the solution u i exp(i ωt) there is always a second
solution (−u i ) exp(−i ωt). This makes it possible to form real combinations from
these roots, which only have a mechanical meaning.
The equation linking the quantities u ik and σikj follows the from (9.23):

σikj , j − ρω2k u ik = 0. (9.24)


248 9 Dynamic Problems of Solid Mechanics

Obviously, due to the homogeneity of the system of equations and boundary


conditions, the required functions included in (9.24) are determined with accuracy
to an arbitrary factor.
Note that Eq. (10.6) can be considered as equations of the static problem of the
theory of elasticity with mass forces −ρω2 u i . Let ω = ωk be any fundamental
frequency. Then u i = u ik is a displacement caused by the action of distributed forces
Pik = −ρω2k u ik . Similarly, forces Pis = −ρω2s u is at frequency ωs cause displacements
u is . By Betty’s theorem, we have:
∫ ∫ ∫ ∫
Pis u ik d V = Pik u is d V or ω2s ρu is u ik d V = ω2k ρu ik u is d V .
V V V V

Since ωs /= ωk , then the last equality is possible if the integral is zero. Therefore,

ρu ik u is d V = 0 (k /= s).
V

This equality expresses the orthogonality property of the fundamental forms


of oscillation. From the condition of orthogonality, it follows, in particular, that
frequencies ωk are always real number.
Since u ik are defined only by the accuracy of an arbitrary constant factor, they can
be normalized arbitrarily. Usually it is accepted:

ρu ik u is d V = δks , (9.25)
V

where δks are Kronecker deltas.


Relations (9.25) express normalization conditions and simultaneously repeat
conditions of orthogonality of fundamental forms of oscillation.
Let’s do some studies of natural oscillation now. We use the superposition principle
(Fourier method). Then we can represent the general expression for displacements
under natural oscillations of the elastic body as follows:

Σ
ui = (Ak sin ωk t + Bk cos ωk t) u ik (xs ), (9.26)
k=1

where Ak , Bk are undefined constants.


Differentiating (9.26) by time and the following is obtained:

Σ
u̇ i = (Ak ωk cos ωk t − Bk ωk sin ωk t) u ik (xs ).
k=1
9.5 Natural and Forced Harmonic Oscillations 249

Equating at t = 0, the values of displacements and velocities are their initial


values u 0i , v0i and we obtain:

Σ ∞
Σ
Bk u ik (xs ) = u 0i (xs ) , Ak ωk u ik (xs ) = v0i (xs ) .
k=1 k=1

We multiply each of these equations by ρu ik and integrate by volume. Due to


(9.25), only one member remains on the left side of each series:
∫ ∫
1
Bk = ρu 0i u ik d V , Ak = ρv0i u ik d V . (9.27)
ωk
V V

It should be noted that the ratios (9.27) do not imply the possibility of decompo-
sition of functions u 0i and v0i into series according to their fundamental forms of
oscillations or fundamental functions u ik . The initial velocity distribution may not
even be continuous at all. Therefore, if we talk about convergence, then we can only
talk about convergence on average.

9.5.2 Forced Oscillations of Elastic Bodies

Let periodic forces act on the body with a circular frequency p. For simplicity, we
will assume that there are no mass forces (F i = 0), and on the body surface, the
external forces are represented in the form: Ri = Ri0 exp(i pt).

Remark Consideration of a more general case of additional difficulties does not


occur.

Let’s assume that the displacements and stresses are also proportional to
exp(ipt).We will also use the designations ui and σij for the amplitudes of
displacements and stresses. From Eq. (9.6), then we obtain the following equation:

σi j , j − ρ p 2 u i = 0. (9.28)

The following boundary conditions shall be performed on the surface S σ :

σi j l j = Ri0 . (9.29)

We present the solution in the form:


/ /
u i = u i + u i0 , σi j = σi j + σi0j , (9.30)
250 9 Dynamic Problems of Solid Mechanics

where u i0 , σi0j are solutions to the static problem of the theory of elasticity. This
solution satisfies the equilibrium equation σi0j , j = 0, Cauchy relations, and boundary
conditions (9.29).
We substitute (9.30) into (9.28). As a result, the first part of the solution in (9.30)
satisfies the following equation of motion:

σi'j , j −ρ p 2 u i' = ρ p 2 u i0 (9.31)

under uniform boundary conditions.


/ /
Present u i , σi j as series:


Σ ∞
Σ
/ /
ui = ak u ik , σi j = ak σikj .
k=1 k=1

Now we substitute these expressions into (9.31) and exclude the stress amplitude
derivatives by using (9.24). As a result, we get:

Σ
ak ρ(ω2k − p 2 )u ik = ρ p 2 u i0 .
k=1

Multiply the obtained equality by u im and integrate by volume. Then we use the
orthonormality condition of eigenfunction (9.25). As a result, we obtain:

am (ω2m −p )= p
2 2
ρu i0 u im d V .
V

From which it follows:



p2
am = 2 ρu i0 u im d V .
ωm − p 2
V

If p = ωm , then the frequency of the perturbing force coincides with one of the
fundamental frequencies of the elastic body, and the corresponding coefficient turns
to infinity over time. This phenomenon is called resonance.
In the case of non-periodic external forces to describe forced oscillations of the
elastic body, the surface load and the solution are presented as decompositions in a
series according to the system of eigenvalue functions. Substituting these series into
equations of motion allows us to obtain equations for determining unknown time
functions.
9.6 Propagation of Shock Waves in Unlimited Elastic Bodies 251

9.6 Propagation of Shock Waves in Unlimited Elastic


Bodies

Waves that have breaks of derivatives from displacements (that is, velocities and
stresses) are called strong breaking waves or shock waves.
The velocity of propagation of the longitudinal wave in the rod is:
/ /
c0 = E ρ. (9.32)

Remark When proving this statement, it is immediately assumed that at the front of
the wave, the velocity and deformation and therefore the stress, change by a jump.

According to the definition, the field of displacements corresponding to the plane


longitudinal wave propagating in the isotropic medium has the form:

u i = Ui (ct − n s xs ). (9.33)

If you substitute (9.33) into the equations of motion (9.9) and solve it, then it
can be established that in an isotropic body, there are two velocities
√ of plane wave
propagation
/ / c 1 and c 2 , having the “classical” expressions c1 = (λ + 2μ)/ρ, c2 =
μ ρ.
As can be seen, the velocity of propagation of the longitudinal wave in the rod
(9.32) is different from the velocities c1 and c2 .
The possibility of propagating shock waves in an unlimited elastic medium with
velocities c1 and c2 requires justification. Let’s show that the longitudinal waves of
strong rupture in an unlimited elastic medium can propagate.
Consider the longitudinal waves propagating along the x1 axis (Fig. 9.2). the
section of medium of the length ct along x1 axis after time interval t will be uniformly
deformed and the rest of medium remains not stressed, i.e., the section (mn), which
is the boundary between the stressed and non-stressed parts of the medium, will
move at a velocity c. We fix a section (pq) with a coordinate x1 . The distance from
the section (pq) to the boundary (mn) is (ct − x1 ). Section with the length (ct − x1 )
uniformly compressed, its deformation is e11 . Therefore, the displacement of the
cross section (pq) from the initial position is u 1 = e11 (ct − x1 ).
Let the wave pass through section 1–1 at time t and through section 2-2 at time
t + dt. The distance among these sections is d x1 = cdt. We differentiate by time
and obtain the velocity of the section movement:
/
v = du 1 dt = e11 c.

For a period of time dt, a force σF acts in the section 1-1 (F is the cross-sectional
area), while the section 2-2 remains not stressed. Therefore, the impulse equals to
σ11 Fdt. At the beginning of time, all the selected parts were at rest, and at the moment
252 9 Dynamic Problems of Solid Mechanics

Fig. 9.2 Longitudinal waves


propagation

t + dt, it is all moving at velocity v. Therefore, there is a change in the momentum:


vρFd x1 = vρFcdt, i.e., σ11 Fdt = vρFcdt, or

σ11 = vρc = e11 ρc2 . (9.34)

Based on the Hooke’s law for isotropic media:

σ11 = (λ + 2μ)e11 + λ(e11 + e22 ).


( )/
From Cauchy ratios ei j = u i, j + u j,i 2 and conditions u 2 = u 3 = 0, it follows
that e11 = e22 = 0, which means:

σ11 = (λ + 2μ)e11 . (9.35)

Substituting (9.35) into (9.34),


√ we obtain the wave propagation velocity in an
unlimited elastic medium: c = (λ + 2μ)/ρ.
Thus, the equations of motion of the elastic medium allow solutions containing
breaks of the first derivatives from displacements.

9.7 Progressive Waves

Progressive waves mean partial solutions of the equations of the dynamic theory of
elasticity, corresponding to waves propagating along a line in the absence of initial
conditions.
Suppose that the studied progressive waves propagate along a line, as which the
axis Ox, x = x1 is selected.
Then the corresponding partial solution of the equations of motion can be taken
in the following form:

u k (t, x1 , x2 , x3 ) = Vk (x2 , x3 )e−iq(x−ct) , (q > 0, c > 0). (9.36)


9.7 Progressive Waves 253

Solutions (9.36) make sense for regions of the infinite size along the Ox axis, for
example, infinite cylinder, in general, of variable section (waveguide), half-space,
plane infinite layer. Included in (9.36), values have the following names: Vk is the
wave amplitude; q is the wave number; ω = qc is the phase frequency; −q(x − ct) is
the wave phase; value L = 2π/q is the wave length (period). In addition, since there
is a point at which the phase keeps a constant value, that is −q(x − ct) = const, so
the value c is called the phase velocity. As with harmonic oscillations, the complex
solution (9.36) should be understood as two processes: its real part changes according
to the law of cosine, imaginary part changes according to the law of sine.
If there are no perturbations (zero mass forces and homogeneous boundary condi-
tions), then the problem of progressive waves is reduced to the problem of eigen-
values. The eigenvalue is the phase velocity c, which in general implicitly determines
the dependence c = c(q) and the corresponding system of eigenfunctions (modes)
Vk (q, x2 , x3 ).
Media (waves) have dispersion if the function c(q) /= const and do not have
dispersion otherwise.
In stationary wave theory, the concept of group velocity cg , introduced by Stokes
and Rayleigh, is used, which is defined as the velocity of the point x = x∗ in which
the phase of the wave is stationary, that is:

d
(−q(x∗ − ct)) = 0. (9.37)
dq

From which it follows:


/
cg = x∗ t = c + qc' (q). (9.38)

From (9.38), it follows that group and phase velocities coincide only with non-
dispersed media.
The importance of information about group velocity is confirmed by the following
qualitative considerations. Consider several (group) waves of the form (9.36) with
close phase frequencies. If at some points, wave phases coincide or are close, then
the amplitudes of individual waves add up modulus. The smallest phase difference
will be provided by (9.37). Therefore, the group velocity defined by formula (9.38) is
the propagation velocity of the maximal perturbation formed by the group of waves.
It can be shown that the concept of progressive waves, as for harmonic waves,
has a direct connection with the complex Fourier transform over the x coordinate.
Namely, if we are looking for a solution in the following form:

u k (t, x, x2 , x3 ) = vk (x − ct, x2 , x3 ), (9.39)

then there is an equality:

vkF (q, x2 , x3 ) = Vk (q, x2 , x3 ), (9.40)


254 9 Dynamic Problems of Solid Mechanics

where the index “F” denotes the image; q is the transformation parameter.
Then the amplitude of the group wave is the Fourier image of the function
vk (x, x2 , x3 ) in (9.40). At the same time, (9.36) is (with accuracy to the constant
factor) a sub-integral expression in the inverse integral of the Fourier transform.
Note that (9.40), there is a special case of auto-model solutions, that is, solutions
that depend on a smaller number of variables than the original problem.
If there is an infinitely distant point in the body, the limitation condition of the
solution specified in Sect. 9.1 must be replaced by the emission condition intro-
duced by Sommerfeld,2 which is the condition for the existence of a single solution
corresponding to the Helmholtz wave equation: Δϕ̃ + k 2 ϕ̃ = Φ , ˜ k > 0, where ϕ̃
and Φ ˜ amplitudes or modes ϕ and Φ .
The emission condition introduced by Sommerfeld has the form (r is the length
of the radius vector; i is an imaginary unit):
• Spatial and one-dimensional problems
( / )
ϕ,r + ikφ = o 1 r , r → ∞,

• Plane problems
( /√ )
ϕ,r + ikϕ = o 1 r , r → ∞.

The concept of progressive waves is widely used when considering waves


propagating over the surface of a semi-space or flat layer.

9.8 Rayleigh Waves

Consider the problem of propagation of a plane progressive wave (9.36) in the direc-
tion of the boundary of the half-plane occupied by a homogeneous linear elastic
isotropic medium (plane-strain state condition). The half-plane boundary is free from
stresses, there are no mass forces and disturbances at infinity.
We introduce such a rectangular Cartesian coordinate system O x yz that the Oz
axis is directed deep into the half-plane and the Ox axis coincides with the boundary
z = 0.
The motion of the medium is described using equations in potentials (9.12) (ϕ =
ϕ(x, z),ψ1 = ψ2 ≡ 0, ψ3 = ψ(x, z)):

∂ 2ϕ ∂ 2ψ ∂2 ∂2
= c12 Δϕ, = c 2
Δψ, Δ = + . (9.41)
∂t 2 ∂t 2 2
∂x2 ∂ z2

Displacements, strains, stresses, and potentials are related by known relationships:

2 Arnold Johannes Wilhelm Sommerfeld, 1868–1951, German theoretical physicist and mathemati-
cian.
9.8 Rayleigh Waves 255

∂ϕ ∂ψ ∂ϕ ∂ψ
u1 = − , u3 = + ,
∂x ∂z ∂z ∂x
( )
∂u 1 1 ∂u 3 ∂u 1 ∂u 3 (9.42)
ε11 = , ε13 = + , ε33 = ,
∂x 2 ∂x ∂z ∂z
σ11 = (λ + 2μ)ε11 + λε 33 , σ13 = 2με 13 , σ33 = (λ + 2μ)ε33 + λε 11 .

The conditions at the boundary of the half-plane and at infinity have the form:

σ13 |z=0 = σ33 |z=0 = 0, ϕ = O(1), ψ = O(1)(z → +∞). (9.43)

To find the solution of the problem (9.41)–(9.43) we will look in the form of a
progressive wave:

ϕ(t, x, z) = Φ (z)E(x, t), ψ(t, x, z) = ψ(z)E(x, t), E(x, t) = e−iq(x−ct) .


(9.44)

Substituting (9.44) into (9.41), we obtain equations regarding the functions Φ (z)
and Ψ(z):
/
Φ '' (z) − q 2 β1 Φ (z) = 0, ψ'' (z) − q 2 β2 ψ(z) = 0, β j = 1 − c2 c2j . (9.45)

With c ≥ c j (β j ≤ 0), solutions to these equations do not satisfy the boundary


conditions at infinity (see 9.43). If c < c j , then β j > 0 (see inequality in (9.12)) and
the solutions, satisfying the condition at infinity, have the form:
∫ / /
Φ (z) = C1 E 1 (z), ψ(z) = C2 E 2 (z), E j (z) = e−qα j z , α j = β j = 1 − c2 c2j ,

where C1 and C2 are arbitrary constants. Therefore:

ϕ(t, x, z) = C1 E 1 (z)E(x, t), ψ(t, x, z) = C2 E 2 (z)E(x, t).

Then from (9.42), we obtain:


• Displacements

u 1 = q[−iC1 E 1 (z) + α2 C2 E 2 (z)]E(x, t),


u 3 = −q[α1 C1 E 1 (z) + iC2 E 2 (z)]E(x, t);

• Strains

ε11 = −q 2 [C1 E 1 (z) + iα2 C2 E 2 (z)]E(x, t),


q2 [ ( ) ]
ε13 = 2i α1 C1 E 1 (z) − 1 + α22 C2 E 2 (z) E(x, t),
2[ ]
ε33 = q 2 α21 C1 E 1 (z) + i α2 C2 E 2 (z) E(x, t),
256 9 Dynamic Problems of Solid Mechanics

c2
θ = −q 2 C1 E 1 (z)E(x, t);
c12

• Stresses
[( ) ]
c2
σ11 = −μq 2
2α1 + 2 C1 E 1 (z) + 2i α2 C2 E 2 (z) E(x, t),
2
c2
[ ( ) ] (9.46)
σ13 = μq 2i α1 C1 E 1 (z) − 1 + α22 C2 E 2 (z) E(x, t),
2
[( ) ]
σ33 = μq 2 1 + α22 C1 E 1 (z) + 2i α2 C2 E 2 (z) E(x, t).

Substituting Eq. (9.46) into the first two boundary conditions (9.43), we obtain a
homogeneous system of linear algebraic equations with respect to constants C1 and
C2 :
( ) ( )
2i α1 C1 − 1 + α22 C2 = 0, 1 + α22 C1 + 2i α2 C2 = 0. (9.47)

The condition for the existence of a non-trivial solution of the system (9.47) is
the equality to zero of its determinants:

( )2 /( / )∫
R(α1 , α2 ) = 1 + α22 − 4α1 α2 = R0 (ξ) = (2 − ξ) − 4 1 − ξ η2 1 − ξ = 0,
2

/ /
ξ = c2 c22 , η2 = c12 c22 .
(9.48)

Multiply the left of this equation by the conjugate expression


/ / ∫
R̃0 (ξ) = (2 − ξ)2 + 4 1 − ξ η2 1 − ξ.

As a result, we get a polynomial of the fourth degree

P42 (ξ) = ξP32 (ξ), P32 (ξ) = ξ3 − 8ξ2 + 8(2 + κ)ξ − 8(1 + κ),

where
/ / /
κ = λ (λ + 2μ) = ν (1 − ν) = 1 − 2 η2 . (9.49)

Since in the real domain R̃0 (ξ) > 0, so (9.48) is equivalent to the following cubic
equation (multiplier ξ rejected because it gives zero root, which corresponds to no
wave):

P32 (ξ) = 0. (9.50)


/
Due to the inequality 0 < ν < 1 2 for real solids, the parameters κ and η change
in the following ranges:
9.8 Rayleigh Waves 257

0 < κ < 1, 2 < η < ∞. (9.51)

Limit case κ = 1 (η = ∞) corresponds to acoustic medium. In this case, the


denominator in (9.48) in equality for ξ turns to zero (C 2 = 0). Therefore, in parallel
with Eq. (9.50), we will consider the equation equivalent to it:
( )
P31 (ς) = ς3 − 4(1 − κ)ς2 + 2(2 + κ)(1 − κ)2 ς − 1 − κ2 (1 − κ)2 = 0,

P32 (η2 ς) = η6 P31 (ς),


(9.52)
R0 (η2 ς) R̃0 (η2 ς) = P42 (η2 ς) = P41 (ς) = η2 ςP31 (ς).

The discriminant of the polynomial P31 (ς) has the form:


( p )3 ( q )2 1 8 8
D3 (κ) = + = Δ3 (κ), p = (3κ − 2), q = (45κ − 11),
3 2 54 3 27
11 3 5
Δ3 (κ) = κ3 + κ2 + κ − .
32 16 32
/
As Δ3 (0) < 0, Δ3 (1) > 0 and Δ3 (κ) > 0, then there is a single real root κ∗ of the
polynomial Δ3 (κ). Moreover, for D3 (κ) we have (values κ∗ , ν∗ and η2∗ are related by
ratios (9.49)):

D3 (κ∗ ) = 0,
)
D3 (κ) < 0 (0 ≤ κ < κ∗ , 0 ≤ ν < ν∗ , 2 ≤ η2 < η∗2 ,
/ )
D3 (κ) > 0 (κ∗ < κ < 1 , ν∗ < ν < 1 2, η∗2 < η2 < +∞ ,
κ∗ ≈ 0, 357003205, ν∗ ≈ 0, 263082064, η∗2 = 3, 1104351.

Thus, polynomials P31 (ς) and P32 (ξ) at 0 ≤ κ < κ∗ have three different real roots.
One polynomial root is real and the other two are complex conjugates if κ∗ < κ < 1.
/ ς1 , ς2 /
Note that, for real roots and ς3 polynomials P31 (ς), the following inequalities
are valid: 0 < ς1 = c2R c12 < 1 η2 (0 ≤ κ < 1), ς2,3 > 1 (0 < κ ≤ κ∗ ).
Thus, there is a single-phase velocity
∫ √
c = c R = ξ R c2 = ς R c1 < c2 , ξ R = ξ1 , ς R = ς1 . (9.53)

The general solution of the system of algebraic equations is represented as follows


(C is arbitrary constant):
( ) ∫
C1 = −2i α2R C, C2 = 1 + α22R C, α2R = 1 − ξ R .

We substitute these constants into (9.46) and obtain displacements and stresses:
258 9 Dynamic Problems of Solid Mechanics
[ ( ) ]
u 1 = Cqα2R −2E 1 (z) + 1 + α22R E 2 (z) E(x, t),
[ ]
( ) 1 + α22R
u 3 = Ciq 1 + α2R
2
E 1 (z) − E 2 (z) E(x, t);
2
[( / ) ( ) ]
σ11 = 2iCμq 2 α2R 2α21R + c2 c22 E 1 (z) − 1 + α22R E 2 (z) E(x, t),
( )2
σ13 = Cμq 2 1 + α22R [E 1 (z) − E 2 (z)]E(x, t),
( ) / /
σ33 = 2iCμq 2 α2R 1 + α22R [−E 1 (z) + E 2 (z)]E(x, t), α1R = 1 − ξ R η2 .
(9.54)

Note that the stress–strain state of the solid is not singly determined, since in
formulas (9.54) there is an arbitrary constant factor C. This is because the original
problem is homogeneous. Due to the linearity of the problem, either the real or
imaginary part of the problem is real, respectively.
The resulting waves are called Rayleigh waves (see Sect. 9.1) and the corre-
sponding phase velocity c R is called the Rayleigh wave velocity. Note that, as follows
from Eq. (9.48), the phase velocity in this case does not depend on the wavenumber
q, therefore, the Rayleigh waves ( are dispersion-free.
) These waves, due to exponen-
tial multipliers with indicators −qα j z , decay very quickly in the half-plane depth.
The decaying velocity is determined
/ by the phase velocity q, that is, the wavelength
L of the x coordinate L = 2π q. The greater the phase velocity (the shorter the
wavelength), the faster the decay occurs.
Therefore, Rayleigh waves are surface waves (their main energy is concentrated at
the boundary of the half-plane). These waves are of great importance in seismology,
since they are observed in earthquakes far from the epicenter, and they are the causes
of destruction of objects on the earth surface. Rayleigh waves also play a large role
in ultrasound, also in fault detection.

9.9 Progressive Waves in a Plane Layer

Rayleigh surface waves can exist in half-space. Consider the problem of propagation
of plane progressive waves in a plane elastic layer of the thickness 2 h. Evaluate
the impact of the presence of the second boundary. We use the setting of the model
problem and the approach to build its solution similar to the previous section. The
geometry of the task and the selected rectangular Cartesian coordinate system are
shown in Fig. 9.3.
The motion of the layer is described using Eq. (9.41) and relations (9.42). Assume
that the layer boundaries are planes z = ±h and they are free (see (9.43)). The
boundary conditions are written as follows:

σ13 |z=±h = σ33 |z=±h = 0. (9.55)


9.9 Progressive Waves in a Plane Layer 259

Fig. 9.3 To the problem of


the propagation of plane
progressive waves in a plane
elastic layer

Just as in the Rayleigh wave problem, we build a solution in the form (9.44) and
as a result we come to Eq. (9.45). The general solution of these equations depends
on the magnitude of the phase velocity: c < c2 , c = c2 , c2 < c < c1 , c = c1 or
c > c1 .
For the examining problem in ranges c < c2 , c2 < c < c1 and c > c1 it is
convenient to write the solutions of Eq. (9.45) in the same form (C k and Dk are
arbitrary constants):
∫ ∫
Φ (z) = C1 sh(qz β1 ) + C2 ch(qz β1 ),
∫ ∫ / (9.56)
Ψ(z) = D1 sh(qz β2 ) + D2 ch(qz β2 ), β j = 1 − c2 c2j .

Remark The results for the bounds of the ranges are obtained using the corre-
sponding limit transitions.

From (9.44) and (9.56), we get the following representations for potentials:
( ∫ ∫ )
ϕ(t, x, z) = C1 sh(qz β1 ) + C2 ch(qz β1 ) E(x, t),
( ∫ ∫ )
ψ(t, x, z) = D1 sh(qz β2 ) + D2 ch(qz β2 ) E(x, t).

Then from (9.42), we find:


{( ∫ ∫ )
u 1 = − q i C1 sh(qz β1 ) + C2 ch(qz β1 )
∫ ( ∫ ∫ )}
+ β2 D1 ch(qz β2 ) + D2 sh(qz β2 ) E(x, t) ,
{∫ ( ∫ ∫ )
u 3 = q β1 C1 ch(qz β1 ) + C2 sh(qz β1 )
( ∫ ∫ )}
− i D1 sh(qz β2 ) + D2 ch(qz β2 ) E(x, t)
{ ∫ ∫
ε11 = −q 2 C1 sh(qz β1 ) + C2 ch(qz β1 )
∫ ( ∫ ∫ )}
− i β2 D1 ch(qz β2 ) + D2 sh(qz β2 ) E(x, t)
q2 { ∫ ( ∫ ∫ )
ε13 = − 2i β1 C1 ch(qz β1 ) + C2 sh(qz β1 )
2
260 9 Dynamic Problems of Solid Mechanics
( ∫ ∫ )}
+ (1 + β2 ) D1 sh(qz β2 ) + D2 ch(qz β2 )) E(x, t)
{ ( ∫ ∫ )
ε33 = q 2 β1 C1 sh(qz β1 ) + C2 ch(qz β1 )
∫ ( ∫ ∫ )}
−i β2 D1 ch(qz β2 ) + D2 sh(qz β2 ) E(x, t)
{ ( ∫ ∫ )
σ11 = μq 2 −(1 + 2β1 − β2 ) C1 sh(qz β1 ) + C2 ch(qz β1 )
∫ ( ∫ ∫ )}
+2i β2 D1 ch(qz β2 ) + D2 sh(qz β2 ) E(x, t)
{ ∫ ( ∫ ∫ )
σ13 = −μq 2 2i β1 C1 ch(qz β1 ) + C2 sh(qz β1 )
( ∫ ∫ )}
+(1 + β2 ) D1 sh(qz β2 ) + D2 ch(qz β2 )) E(x, t)
{ ( ∫ ∫ )
σ33 = q 2 (1 + β2 ) C1 sh(qz β1 ) + C2 ch(qz β1 )
∫ ( ∫ ∫ )}
−2i β2 D1 ch(qz β2 ) + D2 sh(qz β2 ) E(x, t). (9.57)

Substituting Eq. (9.57) into boundary conditions (9.55), after transformations,


we obtain two independent homogeneous systems of linear algebraic equations with
respect to C 2 , D1 and C 1 , D2 :
∫ ∫ ∫
2i β1 C2 sh(qh β1 ) + (1 + β2 )D1 sh(qh β2 ) = 0,
∫ ∫ ∫
(1 + β2 )C2 ch(qh β1 ) + 2i β2 D1 ch(qh β2 ) = 0,
∫ ∫ ∫ (9.58)
2i β1 C1 ch(qh β1 ) + (1 + β2 )D2 ch(qh β2 ) = 0,
∫ ∫ ∫
(1 + β2 )C1 sh(qh β1 ) + 2i β2 D1 sh(qh β2 ) = 0.

As follows from (9.57), with C1 = D2 = 0 the displacements u1 and the stress σ11 ,
σ33 are even functions and the displacements u3 and the stress σ13 are odd functions
along the z coordinate. If C2 = D3 = 0, then on the contrary, u3 and σ13 are even
functions and u1 and σ11 , σ33 are odd functions.
We call these two types of waves symmetric and antisymmetric, respectively, and
due to the linearity of the problem, consider them separately.
1. Symmetric waves. The condition for the existence of a non-trivial solution of
the system (9.46) is the equality to zero of its determinants, which is equivalent to the
transcendent equation (the equation is written at various values of the phase velocity
taking into account the equality thi x = itgx):
9.9 Progressive Waves in a Plane Layer 261
/ /
th(qh 1 − ξ η2 )
(2 − ξ)2
• At c < c2 (ξ < 1) √ = / / ;
th(qh 1 − ξ) √
4 1 − ξ 1 − ξ η2
/ /
( ) th(qh 1 − ξ η2 )
(2 − ξ)2
• At c2 < c < c1 1 < ξ < η2 √ = / / ; (9.59)
tg(qh ξ − 1) √
4 0 − 1 1 − ξ η2
/ /
( ) tg(qh ξ η2 − 1) (2 − ξ)2
• At c > c1 ξ > η2 √ =− / / .
tg(qh ξ − 1) √
4 ξ − 1 ξ η2 − 1

Parameter designations are given in (9.48).


Due to the periodicity of the tangent, Eq. (9.59) have a counting set of solutions
ξ√s1 < ξs2 < ... < ξsk < ... that correspond to phase velocities csk = ςsk c2 , ςsk =
ξsk .
We investigate the behavior of phase velocities in two limit cases: a wavelength
is much longer than the layer thickness (L >> 2h, qh → 0) and a wavelength is
much smaller than the layer thickness (L << 2h, qh → ∞). In this case, we will
use the following properties of elementary functions:
/ /
tgx ∼ x ∼ x + x 3 3, thx ∼ x ∼ x − x 3 3, x → 0, thx ∼ 1, x → +∞.
(9.60)

In the third Eq. (9.59), we keep the terms of the first order ) at qh → 0.
( of /small
As a result, we get (equivalent signs replace equalities): 4 1 − ξ η2 = (2 − ξ)2 .
Hence, for the first waveform we have that at qh → 0 (see inequalities (9.51)):

( / )/
ξs1 → ξ∗ = 2(1 + κ), c → c∗ = 2 c2 c1 c12 − c22 , 1 < 2(1 + κ) ≤ η2 .
(9.61)

The use of decompositions from (9.60), taking into account the terms of the third
order, leads this equation at qh → 0 to the form:
⎧ 2 2 [ ( )] ( )⎧
q h 2 q 2h2 4 2q 2 h 2
ξ ξ 1− 5 − 2 ξ − 2(1 + κ) 1 − = 0.
3 3 η 3

Since the coefficient at the highest degree tends to zero, the root of the equation
tends to infinity.
A similar relationship analysis (9.59) results in the same result that corresponds
to the range of change ξ corresponding to the third of the Eq. (9.59). Since the first
form is true (9.61), so this conclusion refers to the higher forms ξsk → ∞, csk → ∞
at qh → 0 (k ≥ 2). Considering the last equivalence ratio in (9.60), from (9.59),
when qh → 0, we obtain the equation:
262 9 Dynamic Problems of Solid Mechanics

(2 − ξ)2
√ / / = 1.
4 1 − ξ) 1 − ξ η2

This equation is equivalent to Eq. (9.47). Therefore, the phase velocity of the first
waveform tends to the Rayleigh wave velocity (see (9.53)): ξs1 → ξ R , c → c R . That
is, in this case, the boundaries of the layer z = ±h have little influence on each other
and behave like the boundaries of the half-space. Thus, at the layer boundaries there
are mainly waves close to Rayleigh waves.
From the relations (9.59), it follows that waves of this type are dispersed. The
group speed according to (9.39) will be:

cg ξ' (qh)
ςg = = ς(qh) + qh, (9.62)
c2 2ς(qh)

where ξ' (qh) is a derivative of the function ξ(qh), which in turn is implicitly given
by Eq. (9.59).
It can be shown that the group velocity for the first form has the same limit values
as the phase velocity and for higher forms, it tends to zero at qh → 0.
To determine the waveform, put in (9.58)
∫ ∫ ∫
C2 = C(1 + β2 )sh(qh β2 ), D1 = −2iC β1 sh(qh β1 ),

where C is an arbitrary constant.


Substituting these equations into (9.57), we obtain the waveforms of displace-
ments and stresses (C 1 = D2 = 0):
{ ∫ ∫ ∫ ∫ ∫ ∫ }
u 1 = iqC −(1 + β2 )sh(qh β2 )ch(qz β1 )+ 2 β1 β2 ch(qz β2 )sh(qz β1 ) E(x, t),
∫ { ∫ ∫ ∫ ∫ }
u 3 = qC β1 (1 + β2 )sh(qh β2 )sh(qz β1 )− 2sh(qh β1 )sh(qz β2 ) E(x, t),
{ ∫ ∫
σ11 = μq 2 C −(1 + 2β1 − β2 )(1 + β2 )sh(qh β2 )ch(qz β1 )
∫ ∫ ∫ ∫ }
+4 β1 β2 sh(qh β1 )ch(qz β2 ) E(x, t),
∫ { ∫ ∫
σ13 = 2i μq 2 C β1 (1 + β2 ) - sh(qh β2 )sh(qz β1 )
∫ ∫ }
+sh(qh β1 )sh(qz β2 ) E(x, t),
{ ∫ ∫
σ33 = q 2 C (1 + β2 )2 sh(qh β2 )ch(qz β1 )
{ ∫ ∫
− σ33 = q 2 C (1 + β2 )2 sh(qh β2 )ch(qz β1 )
∫ ∫ ∫ ∫ }
−4 β1 β2 sh(qh β1 )ch(qz β2 ) E(x, t).
9.10 Semi-plane Under Action of Moving Surface Force 263

2. Antisymmetric waves. Similar to symmetric waves, from the condition of the


existence of a non-trivial solution of the system (9.58), we obtain a transcendent
equation with respect to a dimensionless phase velocity:
/ / / /

th(qh 1 − ξ η2 ) 4 1 − ξ) 1 − ξ η2
• If c < c2 (ξ < 1) then √ = ;
th(qh 1 − ξ)) (2 − ξ)2
/ / / /

th(qh 1 − ξ η2 ) 4 ξ − 1 1 − ξ η2
(9.63)
• If c2 < c < c1 (1 < ξ < η) then √ =− ;
tg(qh ξ − 1) (2 − ξ)2
/ / / /

( ) tg(qh ξ η2 − 1) 4 ξ − 1 ξ η2 − 1
2
• If c > c1 ξ > η then √ =− .
tg(qh ξ − 1) (2 − ξ)2

Equation (9.63) have a counting set of solutions ξα1 <√ξα2 < ... < ξαk < ..., to
which the phase rates correspond to cak = ςak c2 , ςak = ξak .
A study of the asymptotic behavior of phase and group velocities (see (9.62))
shows that for the first form ξa1 → 0, ξga1 → 0 at qh → 0 and ξa1 → ξ R , ξga1 →
ξ R at qh → ∞; for higher forms ξsk → ∞, ξgsk → 0 at gh → 0 (k ≥ 2).
To determine the shape of antisymmetric oscillations, you can put it in (9.58) (D
is an arbitrary constant):
∫ ∫ ∫
C1 = D(1 + β2 )ch(qh β2 ), D2 = −2i D β1 ch(qh β1 ).

Then it is necessary to substitute these equalities into (9.57), considering that C 2


= 0 and D1 = 0.

9.10 Semi-plane Under Action of Moving Surface Force

Consider another practically important problem about the movement of force P with
a constant velocity v in the direction of the axis Ox, which is normal to the boundary
z = 0 of the half-plane (Fig. 9.4).
Motion of elastic medium is described using ratios (9.41)–(9.43). Boundary
conditions on the half-plane boundary have the form:

σ13 |z=0 = 0, σ33 |z=0 = −PΔ(x − vt), (9.64)

where δ(x) is the Dirac delta function.

Remark We formulate infinity conditions later.

We assume that the solution of the problem is functions from x − vt and z.


Therefore, we move to the new mobile coordinate system x1 = x − vt, z 1 = z.
In new variables, Eq. (9.41) take the following form:
264 9 Dynamic Problems of Solid Mechanics

Fig. 9.4 To the problem of


motion with a constant speed
of force P normal to the
boundary of the half-plane

∂ 2ϕ ∂ 2ϕ ∂ 2ψ ∂ 2ψ v
β1 + = 0, β 2 + 2 = 0, β j = 1 − M 2j , M j = , (9.65)
∂ x1
2 ∂z 2 ∂ x1
2 ∂z cj

where M j are Mach numbers3 ).


The ratios (9.42) are retained. Only in (9.42) and (9.43) the x coordinate is replaced
by a variable x 1 . Boundary conditions (9.64) move to the following:

σ13 |z=0 = 0, σ33 |z=0 = −Pδ(x1 ). (9.66)

Depending on the ratio of the velocity motion v, the acting force P and the propa-
gation speeds c1 , c2 (Mach numbers, the coefficient β j ), the Eq. (9.65) have different
types:
• Both equations are elliptical, if v < c2 < c1 (M1 < M2 < 1,0 < β2 < β1 );
• The first equation is elliptical and the second is hyperbolic, if c2 < v < c1
(1 < M1 < M2 ,β2 < β1 < 0);
• Both equations are hyperbolic, if c2 < c1 < v (1 < M1 < M2 ,β2 < β1 < 0).
These ranges have names: the first of them is subsonic, the second is transonic,
and the third is supersonic.
We use the complex Fourier transform to solve the problem by the coordinate x 1
(index “F” indicates a transform, q is the transformation parameter).
In the transformation space of Eq. (9.65), the relations (9.42) and the boundary
conditions (9.66), take the following form:

3 Ernst Mach (1838–1916), Austrian physicist, mechanic.


9.10 Semi-plane Under Action of Moving Surface Force 265

∂ 2ϕF ∂ 2ψF
−q 2 β1 ϕ F + = 0, −q 2
β 2 ψ F
+ = 0,
∂z 2 ∂z 2
∂ψ F ∂ϕ F
u 1F = −iqϕ F − , u 3F = − iqψ F ,
∂z ∂z (9.67)
( )
1 ∂u F ∂u F
ε11
F
= −iqu 1F , ε13F
= −iqu 3F + 1 , ε33 F
= 3,
2 ∂z ∂z
F|
| F|
|
σ 13 z=0 = 0, σ 33 z=0 = −P.

In the Hooke’s law (9.42), the index “F” must be added to the designations of the
tensor component.
Next, consider separately all three ranges in which the velocity v may be located.
1. Subsonic velocity v < c2 (M1 < M2 < 1,0 < β2 < β1 ).
Equation (9.67) in this case take the form:

∂ 2ϕF ∫ ∂ 2ψF ∫
− q 2 α21 ϕ F = 0, α1 = β1 , − q 2 α22 ψ F = 0, α2 = β2 . (9.68)
∂z 2 ∂z 2

The corresponding infinity conditions are the attenuation conditions of the


solutions of these equations:

ϕ F = O(1), ψ F = O(1) (z → +∞). (9.69)

Solutions of Eq. (9.68) satisfying the conditions (9.69) have the form:

ϕ F (q, z) = C1 E 1 (z), E 1 (z) = e−α1 |q|z , ψ F (q, z) = C2 E 2 (z), E 2 (z) = e−α2 |q|z ,
(9.70)

where C 1 and C 2 are arbitrary constants.


From (9.67) and (9.42), we find displacements:

u 1F = −iqC1 E 1 (z) + α2 |q|C2 E 2 (z), u 3F = −α1 |q|C1 E 1 (z) − iqC2 E 2 (z);


ε11
F
= −q[qC1 E 1 (z) + i|q|α2 C2 E 2 (z)],
q[ ( ) ]
ε13
F
= 2i|q|α1 C1 E 1 (z) − q 1 + α22 C2 E 2 (z) ,
2[ ]
ε33
F
= q qα21 C1 E 1 (z) + i|q|α2 C2 E 2 (z) ; (9.71)
[ ( ) ]
σ11
F
= −μq q 2α21 + M22 C1 E 1 (z) + 2i|q|α2 C2 E 2 (z) ,
[ ( ) ]
σ13
F
= μq 2i|q|α1 C1 E 1 (z) − q 1 + α22 C2 E 2 (z) ,
[ ( ) ]
σ33
F
= μq q 1 + α22 C1 E 1 (z) + 2i|q|α2 C2 E 2 (z) .

Note that when q > 0 and v = c in accordance with (9.40), the formulas (9.71)
coincide with the equalities (9.46)–(9.47).
266 9 Dynamic Problems of Solid Mechanics

Substituting stresses from (9.71) into boundary conditions in (9.67), we obtain a


system of linear algebraic equations with respect to constants C 1 and C 2 :
( ) ( ) /
2i|q|α1 C1 − q 1 + α22 C2 = 0, q 1 + α22 C1 + 2i|q|α2 C2 = −P (μq). (9.72)

The determinant of this system equals to q 2 R(α1 , α2 ), that is, it is proportional to


the determinant (9.47) at v = c. Its solution has the form:
( )
P K u 1 + α22 2i P K u α1 signq 1
C1 = − , C2 = − , Ku = .
μq 2 μq 2 R(α1 , α2 )

Substituting these constants into (9.71), we find an image of the displacements


and stresses:
i P K u [( ) ]
u 1F = 1 + α22 E 1 (z) − 2α1 α2 E 2 (z) ,
μq
P K u α1 [( ) ]
u 3F = 1 + α22 E 1 (z) − 2E 2 (z) ,
μ|q|
[( )( ) ]
σ11 = P K u 1 + α22 2α21 + M22 E 1 (z) − 4α1 α2 E 2 (z) ,
F
( )
σ13
F
= −2i P K u α1 1 + α22 [E 1 (z) − E 2 (z)]signq,
[( )2 ]
σ33
F
= −P K u 1 + α22 E 1 (z) − 4α1 α2 E 2 (z) .

Using the Fourier transform properties, we finally find displacements and stresses
from these relations:
[ ]
P Ku ( ) x1 x1
u 1 (x1 , z) = 1 + α2 arctg
2
− 2α1 α2 arctg ,
πμ α1 z α2 z
[ ]
P K u α1 ( ) / 2 /
u3 = C M2 − 1 + α2 ln x1 + α1 z + 2 ln x1 + α2 z ,
2 2 2 2 2 2 2
πμ
[ ]
P K u α1 z ( 2 ) 1 + α22 4α22
σ11 = 2α1 + M22 2 − ,
π x1 + α21 z 2 x12 + α22 z 2
( )( )
2P K u α1 1 + α22 M22 − M12 x1 z 2
σ13 = ( )( ) ,
π x12 + α21 z 2 x12 + α22 z 2
[( )2 ]
P K u α1 z 1 + α22 4α22
σ33 = − − 2 .
π x12 + α21 z 2 x1 + α22 z 2

It is obvious if the velocity


√ of the force√ movement coincides with the velocity of
Rayleigh waves cR (M2 = ξ R , M1 = ζ R , see (9.53)), then all components of the
solid stress–strain state strive for infinity. That is, the solution to the problem does
not exist (there is no solution to the system of Eq. (9.72)).
9.10 Semi-plane Under Action of Moving Surface Force 267
/
The limit value of the force velocity v = c2 (M2 = 1,M1 = 1 η, α2 = 0,
/ /
α1 = (1 + κ) 2, K u = 1) is not critical, that is, the solution to the problem exists
and has the form:

P x1 2
u 1 (x1 , z) = arctg √ ,
πμ z 1+κ
⎡ / ⎤

P 1 + κ⎣ 2x12 + (1 + κ)z 2
u3 = √ C − ln √ ⎦,
πμ 2 x12 2
√ √
Pz(2 + κ) 2(1 + κ) P z 2 (1 + κ) 2(1 + κ)
σ11 = [ 2 ] , σ13 = − [ 2 ] ,
π 2x1 + (1 + κ)z 2 π 2x1 + (1 + κ)z 2 x1

Pz 2(1 + κ)
σ33 = − [ 2 ]. (9.73)
π 2x1 + (1 + κ)z 2

Analysis of formulas (9.73) shows that stresses are zeros at infinity, tangent
displacements u1 are bounded in the vicinity of a point at infinity x1 = ∞, on
the surface of the half-plane they are bounded but have a first kind of break. Normal
displacements u3 have a logarithmic feature at the beginning of the moving coordinate
system and at x1 = ∞:

P Ku ( ) P K u α1 M22
u 1 |z=0 = 1 + α22 − 2α1 α2 signx1 , u 3 |z=0 = (C + ln|x1 |);
πμ πμ
P Ku ( ) P K u α1 M22
lim u 1 = ± 1 + α22 − 2α1 α2 , u 3 ∼ ln|x1 |(x1 → ±∞).
x1 →±∞ πμ πμ

Note that the presence of features in displacements does not contradict the medium
continuity hypothesis, since an idealized concentrated load is considered. In the case
of distributed load, these features are not present.
2. Transonic speed c2 < v < c1 (M1 < 1 < M2 ,β2 < 0 < β1 ).
Equation (9.68) together with infinity conditions (9.69) are kept. Also in this
case, the corresponding solution is retained (9.70). The second equation from (9.67)
is written as follows:

∂ 2ψF ∫
+ q 2 λ22 ψ F = 0, λ2 = −β2 . (9.74)
∂z 2

For this equation, the infinity condition (one-dimensional problem) is the radiation
condition (9.41):
( )
∂ψ F 1
+ iqλ2 ψ F = o (z → +∞). (9.75)
∂z z
268 9 Dynamic Problems of Solid Mechanics

The solution of Eq. (9.74) satisfying condition (9.75) is as follows:

ψ F (q, z) = C2 Ẽ 2 (z), Ẽ 2 (z) = e−iλ2 qz .

Similarly, in Eq. (9.71) we obtain the representations of transformants of displace-


ments, deformations, stresses, and a system of linear algebraic equations relative to
constants C 1 and C 2 :

u 1F = −iqC1 E 1 (z) + i λ2 qC2 Ẽ 2 (z), u 3F = −α1 |q|C1 E 1 (z) − iqC2 Ẽ 2 (z),


[ ] [ ]
ε11
F
= q 2 −C1 E 1 (z) + λ2 C2 Ẽ 2 (z) , ε33
F
= q 2 α21 C1 E 1 (z) − λ2 C2 Ẽ 2 (z) ,
q[ ( ) ]
ε13
F
= 2i|q|α1 C1 E 1 (z) − q 1 − λ22 C2 Ẽ 2 (z) ,
2 [ ]
( )
σ11 = μq 2 − 2α21 + M22 C1 E 1 (z) + 2λ2 C2 Ẽ 2 (z) ,
F

[ ( ) ]
σ13
F
= μq 2i|q|α1 C1 E 1 (z) − q 1 − λ22 C2 Ẽ 2 (z) ,
[( ) ]
σ33
F
= μq 2 1 − λ22 C1 E 1 (z) − 2λ2 C2 Ẽ 2 (z) .
( ) ( ) /( )
2i|q|α1 C1 − q 1 − λ22 C2 = 0, 1 − λ22 C1 − 2λ2 C2 = −P μq 2 . (9.76)

The solution of the system of Eq. (9.76) has the following form (the polynomial
P31 (ς ) is defined in (9.52)):
( )
P K t 1 − λ22 [( )2 ]
C1 = − 1 − λ22 + 4i α1 λ2 signq ,
μq 2

2i P K t α1 signq [( )
2 2
]
C2 = − 1 − λ + 4i α 1 λ2 signq ,
μq 2 2
( )−1
K t = η2 M12 P31 (M12 ) . (9.77)

Substituting these constants in (9.76) we obtain images of displacements and


stresses:
P K t i [( )2 ] [( ) ]
u 1F = 1 − λ22 + 4i α1 λ2 signq 1 − λ22 E 1 (z) − i2α1 λ2 Ẽ 2 (z)signq ,
μq
P K t α1 [( )2 ] [( ) ]
u 3F = 1 − λ22 + 4i α1 λ2 signq 1 − λ22 E 1 (z) − 2 Ẽ 2 (z) ,
μ|q|
( )[( )2 ]
σ11 = P K t 1 − λ22 1 − λ22 + 4i α1 λ2 signq
F

[( )( ) ]
2α21 + M22 1 − λ22 E 1 (z) − 4iα1 λ2 Ẽ 2 (z)signq
( )[( )2 ][ ]
σ13
F
= −2i P K t α1 1 − λ22 1 − λ22 + 4i α1 λ2 signq E 1 (z) − Ẽ 2 (z) signq,
9.10 Semi-plane Under Action of Moving Surface Force 269
[( )2 ] [( )2 ]
σ33
F
= −P K t 1 − λ22 + 4i α1 λ2 signq 1 − λ22 E 1 (z) − 4i α1 λ2 Ẽ 2 (z)signq .
(9.78)

From (9.78), using the Fourier transform properties, we find displacements and
stresses:
[ ) /
P Kt ( )3 x1 (
u 1 (x1 , z) = 1 − λ22 arctg + 4α1 λ2 1 − λ22 ln x12 + α21 z 2 −
πμ α1 z
( )2 ( ) ]
−2α1 λ2 1 − λ2 ln|x1 + λ2 z| + 4α21 λ22 sign(x1 + λ2 z) + 2α1 λ2 1 − λ42 C ,
2

P K t α1 ( )3 / ( ) x1
u 3 (x1 , z) = [− 1 − λ22 ln x12 + α21 z 2 + 4α1 λ2 1 − λ22 arctg +
πμ α1 z
( )2 ( )( ) ]
+2 1 − λ22 ln|x1 + λ2 z| − 4α1 λ2 sign(x1 + λ2 z) + 1 − λ22 1 − λ42 C ,
⎡ ( )2
( )( ) 1 − λ22 z + 4λ2 x1

σ11 (x1 , z) = P K t α1 ⎣ 2α21 + M22 1 − λ22 ( )
π x12 + α21 z 2
( )2 ⎤
4λ2 1 − λ22

− + 16α1 λ22 δ(x1 + λ2 z)⎦,
π(x1 + λ2 z)
⎧( )( )
( )⎨ 1 − λ22 λ2 x1 − α21 z z
σ13 (x1 , z) = −2P K t α1 1 − λ22 ( )
⎩ π x 2 + α2 z 2 (x + λ z)
1 1 1 2
⎡ ⎤⎫
α1 z ⎬
+4α1 λ2 ⎣δ(x1 + λ2 z) − ( )⎦ ,
π x 2 + α2 z 2 ⎭
1 1
⎡ ( )2
( )2 1 − λ22 z + 4λ2 x1

σ33 (x1 , z) = −P K t α1 ⎣ 1 − λ22 ( )
π x12 + α21 z 2
( )2 ⎤
4λ2 1 − λ22

− + 16α1 λ22 δ(x1 + λ2 z)⎦. (9.79)
π(x1 + λ2 z)

In the transonic case, the argument ς = M12 of polynomial P31 (ς) (see (9.77)) is in
the range η−2 < ς < 1. There are no valid polynomial roots in this range. Therefore,
the solution of the problem, unlike the subsonic case, is determined at any loading
speed from the specified range.
The limit values of the force speed corresponding to the ends of the range are
also not critical in the transonic case. Therefore, displacements and stresses coincide
with corresponding values defined by/formulas (9.73) for subsonic case if v = c2
/ /
(M1 = 1 η, M2 = 1, λ2 = 0, α1 = (1 + κ) 2, K t = 1).
270 9 Dynamic Problems of Solid Mechanics
( ( )2 )
If v = c1 M2 = η, M1 = 1, λ22 = η2 − 1, α1 = 0, K t−1 = η2 − 2 , then
as follows from (9.79), we are dealing with one-dimensional deformation:
( / )
u 1 = P 2λ signx1 , u 3 ≡ 0;
( / )
σ11 = − P κ δ(x1 ), σ13 ≡ 0, σ33 = −Pδ(x1 ). (9.80)

When obtaining the first and third equations in (9.80), the following property of
the delta-set of functions f (x1 , z) at z → +0 was used:
α1 z
lim f (x1 , z) = δ(x1 ), f (x1 , z) = . (9.81)
z→+0 π(x12 + α12 z 2 )

Remark The same result can be obtained, if you perform the limit transition in
Eq. (9.78), and then the inverse Fourier transform.

It follows from (9.79) that displacements have a logarithmic feature at the half-
plane boundary and at x1 → ∞:

P K t {[ π ( )3 ] ( ) }
u 1 |z=0 = 1 − λ22 + 4α21 λ22 signx1 + 2α1 λ2 1 − λ42 (ln|x1 | + C) ,
πμ 2
P K t α1 {( )( ) [ ( ) ] }
u 3 |z=0 = 1 − λ22 1 − λ42 (ln|x1 | + C) + 2α1 λ2 π 1 − λ22 − 2 signx1 ,
πμ
2P K t ( ) P K t α1 ( )2 ( )
u1 ∼ α1 λ2 1 − λ42 ln|x1 |, u 3 ∼ 1 − λ22 1 − λ42 ln|x1 | (x1 → ∞).
πμ πμ

From (9.79), it follows that stresses have a singular feature x1 + λ2 z = 0 in the


front of the Mach wave.
The boundary conditions in (9.67) are fulfilled, since the limit equality (9.81)
is true. For the same reason, the stresses at the half-plane boundary are defined as
follows:
⎧ ( )( )α λ (( )( )3 ) ⎧
1 2
σ11 (x1 , z) = P K t 8 α21 + λ22 1 − λ22 + 2α21 + M22 1 − λ22 + 16α21 λ22 δ(x1 ) .
πx1

3. Supersonic speed c2 < c1 < v (1 < M1 < M2 ,β2 < β1 < 0).

In this case, the solution can also be constructed using the Fourier transform. In
this case, the potential image ϕ must satisfy
√ Eq. (9.74) and condition (9.75), in which
λ2 must be replaced by a value λ1 = −β1 .
However, since in this case both equations in (9.65) are hyperbolic, it is more
convenient to use a different approach based on the functional-invariant solutions of
D’Alembert:

ϕ(x1 , z) = f (x1 + λ1 z), ψ(x1 , z) = g(x1 + λ2 z),


9.10 Semi-plane Under Action of Moving Surface Force 271

where f (x) and g(x) are arbitrary, required number of times of differentiable
functions.
Substituting these representations in (9.42) we find (x = x1 ):

u 1 = f ' (x1 + λ1 z) − λ2 g ' (x1 + λ2 z), u 3 = λ1 f ' (x1 + λ1 z) + g ' (x1 + λ2 z),
ε11 = f '' (x1 + λ1 z) − λ2 g '' (x1 + λ2 z),
1[ ( ) ]
ε13 = 2λ1 f '' (x1 + λ1 z) − λ22 − 1 g '' (x1 + λ2 z) ,
2
ε33 = λ21 f '' (x1 + λ1 z) + λ2 g '' (x1 + λ2 z),
[( ) ]
σ11 = μ M22 − 2λ21 f '' (x1 + λ1 z) − 2λ2 g '' (x1 + λ2 z) ,
[ ( ) ]
σ13 = μ 2λ1 f '' (x1 + λ1 z) − λ22 − 1 g '' (x1 + λ2 z) ,
[( ) ]
σ33 = μ λ22 − 1 f '' (x1 + λ1 z) + 2λ2 g '' (x1 + λ2 z) . (9.82)

Substituting the last two equations from (9.82) into boundary conditions (9.66),
we get a system of linear algebraic equations with respect to f '' (x1 ) and g '' (x1 ):
( )
2λ1 f '' (x1 ) − λ22 − 1 g '' (x1 ) = 0,
( 2 ) ( / )
λ2 − 1 f '' (x1 ) + 2λ2 g '' (x1 ) = − P μ δ(x1 ).

Its solution has the form (see (9.47)):

P Ks ( 2 ) 2P K s
f '' (x1 ) = − λ2 − 1 δ(x1 ), g '' (x1 ) = − λ1 δ(x1 ),
μ μ
(9.83)
Ks =
1 ˜ , λ ) = R(iλ , iλ ) = (λ2 − 1)2 + 4λ λ
, R̃(λ1 2 1 2 1 2
˜
R̃(λ1 , λ2 )
2

By integrating these equations, we find the first derivatives of the functions:

P Ks ( 2 ) 2P K s
f ' (x1 ) = − λ2 − 1 H (x1 ), g ' (x1 ) = − λ1 H (x1 ).
μ μ

Substituting these expressions for derivatives into (9.82) leads to final formulas
for displacements and stresses:

P Ks [ ( 2 ) ]
u1 = − λ2 − 1 H (x1 + λ1 z) + 2λ1 λ2 H (x1 + λ2 z) ,
μ
P K s λ1 [( 2 ) ]
u3 = − λ2 − 1 H (x1 + λ1 z) + 2H (x1 + λ2 z) ;
μ
[ ( )( ) ] (9.84)
σ11 = P K s − M22 − 2λ21 λ22 − 1 δ(x1 + λ1 z) + 4λ1 λ2 δ(x1 + λ2 z) ,
( )
σ13 = 2P K s λ1 λ22 − 1 [δ(x1 + λ2 z) − δ(x1 + λ1 z)],
[( )2 ]
σ33 = −P K s λ22 − 1 δ(x1 + λ1 z) + 4λ1 λ2 δ(x1 + λ2 z) ,
272 9 Dynamic Problems of Solid Mechanics

where H (x) is Heaviside step function.


˜ , λ ) (see (9.83)) in the supersonic case is strictly positive,
Since the function R̃(λ 1 2
the solution to the problem is determined at any loading velocities from the specified
range.
( At the limit value of the supersonic force velocity v = c1
−1
( 2 )2 )
M2 = η, M1 = 1, λ2 = η − 1, λ1 = 0, K s = η − 2 , the stresses and
2 2

normal displacements coincide with the corresponding values determined by the


formulas (9.80) for the transonic case. Only tangent displacements differ:
( / )
u 1 = P λ H (x1 ).

This is due to the hyperbolic type of both Eq. (9.65) at supersonic velocity.
It follows from (9.84) that by the fronts x1 + λ1 z = 0 and x1 + λ2 z = 0 of Mach
waves, the half-plane is divided into three regions in each of which the displacements
are constant and they have breaks of the first kind on the fronts:

• At x1 < −λ2 z, u 1 = u 3 ≡ 0;
2K s λ1 λ2 2K s λ1
• At − λ2 z < x1 < −λ1 z, u 1 = P , u 3 = −P ;
μ μ
Ks ( 2 ) P K s λ1 2
• At x1 > −λ1 z, u 1 = −P λ2 − 1 − 2λ1 λ2 , u 3 = − M2 .
μ μ

Formulas in (9.84) show that stresses have a singular feature on the selected fronts.

9.11 Model Problems on Perturbation Propagation


in Elastic Space

Consider model problems about the propagation of perturbations in an elastic homo-


geneous isotropic space, taking the equations of motion relative to the displacement
x , t):
vector as the original equations u k (→

( ) ∂θ
ü k = c12 − c22 + c22 δu k + Fk , x→ = (x1 , x2 , x3 ) ∈ R 3 , t > 0, (9.85)
∂ xk

where the points denote the time differentiation operation t; Δ is the Laplacian;
/ / expansion-compression and shear wave speeds: c1 =
c1 and c2 are, as before,

(λ + 2μ)/ρ, c2 = μ ρ.
To the system of Eq. (9.85) we add initial conditions to close the problem:

u k |t=0 = ϕk (x1 , x2 , x3 ), u̇ k |t=0 = ψk (x1 , x2 , x3 ), (x1 , x2 , x3 ) ∈ R 3 .


9.11 Model Problems on Perturbation Propagation in Elastic Space 273

The solution of model problems is based on the use of the Green tensor concept.
Let’s give some information about fundamental solutions and the Green tensor of
non-stationary processes.
The Green tensor (influence function) G km (→x , t) of the wave equation is the
solution to the following boundary problems [Gorshkov, A.G. Waves in contin-
uous media: coursebook / A.G. Gorshkov, A.L. Medvedskiy, L.N. Rabinskiy, D.V.
Tarlakovskiy. – Moscow. Pub by “FIZMATLIT”, 2004. 472 p. (in Russian)]:

( )∂ θ
m
G̈ km = c12 − c22 + c22 δG km + δkm δ(t)δ(x1 , x2 , x3 ), (x1 , x2 , x3 ) ∈ R 3 , t > 0,
∂ xk
|
G km |t=0 = 0, Ġ km |t=0 = 0.

1 xk xm 1( xk xm )
G km = δ(c 1 t − r ) + δ km − δ(c2 t − r )
4πr c1r 2 c2 r2
where ( ) ⎧
t 3xk xm
− 2 δkm − [H (c1 t − r ) − H (c 2 t − r )] .
r r2
If the Green function is known, you can define displacements based on the
following integral representations:

x , t) = G km (→
u k (→ x , t) ∗ ∗Fm (→
x , t) + Ġ km (→
x , t) ∗ ϕm (→
x ) + G km (→
x , t) ∗ ψm (→
x ),
(9.86)

where the sign “* ” denotes an operation of function convolution.


In (9.86), the convolution in the first term is calculated both by spatial coordinates
and by time, and in the rest – only by spatial coordinates.
The deformation G εklm and the volumetric expansion factor θm corresponding to
the influence function G km are defined as follows:
( ) ( )
1 ∂G lm ∂G km ∂G 1m ∂G 2m ∂G 3m
G εklm = + ; θm = G ε11m + G ε22m + G ε33m = + .
2 ∂ xk ∂ xl ∂ x1 ∂ x2 ∂ x3

It should be noted that if the Green tensor G km (→ x , t) is a symmetric tensor of the


second rank, then G εklm (→
x , t) is a tensor of the third rank.
Representations for stresses with known tensors G km (→ x , t) and G εklm (→
x , t) have
the form:

x , t) = G σklm (→
σkl (→ x , t) ∗ ∗Fm (→ σ
x , t) + Ġ lkm x ) + G σklm (→
x , t) ∗ ϕm (→
(→ x , t) ∗ ψm (→
x ),
(9.87)

where G σklm (→
x , t) is the influence function for stresses. They are components of the
third-rank tensor and they are defined as follows:

G σklm (→
x , t) = λ θm δkl + 2μG εklm . (9.88)
274 9 Dynamic Problems of Solid Mechanics

By performing the procedure for obtaining influence function G εklm , we get:



1 x k xl x m '
G εklm =− 2 δ (c1 t − r )
4π r c1r 2
( )
1 6xk xl xm
− xk δlm + xl δkm + xm δkl − δ(c1 t − r )
c1 r r2
( )
1 xk δlm + xl δkm x k xl x m '
+ − δ (c2 t − r )
c2 2 r2
( )
1 xk δlm + xl δkm 6xk xl xm
+ 3 + xm δkl − δ(c2 t − r )
c2 r 2 r2
( ) ⎧
3t 5xk xl xm
− 3 xk δlm + xl δkm + xm δkl − [H (c1 t − r ) − H (c2 t − r )] .
r r2
(9.89)

The influence function for stresses G σklm (→


x , t) is determined by simply substituting
Eq. (9.89) into (9.88).
Based on the above formulas, we will construct several concrete types of influence
functions.

Example 1 Find displacements in elastic space under the action of external forces
F1 = δ(t)δ(x1 ), F2 = F3 ≡ 0 and zero initial conditions.

Displacements in this case are defined as:


¨
x , t) = G kl (→
u k (→ x , t) ∗ ∗δ(t)δ(x1 ) = x , t)d x2 d x3 .
G k1 (→ (9.90)
R2

The corresponding influence function for displacements is:


⎧ 2 ( )
1 x1 1 x12
G 11 = δ(c1 t − r ) + 1 − 2 δ(c2 t − r )
4π r c1r 2 c2 r
( 2 ) ⎧
t 3x1
− 1 − 2 [H (c1 t − r ) − H (c2 t − r )] ;
r2 r

x1 xk 1 1
G k1 = 3
δ(c1 t − r ) − δ(c2 t − r )
4π r c1 c2

3t
+ [H (c1 t − r ) − H (c2 t − r )] , (k = 2, 3). (9.91)
r2

Since the function G k1 is odd in coordinate xk , if k = 2, 3, so the corresponding


integrals in (9.90) are zeros and, therefore, at a specified type of disturbance u 2 =
u 3 ≡ 0. Expressions for displacements u 1 , based on (9.90) and (9.91), take the form:
9.11 Model Problems on Perturbation Propagation in Elastic Space 275
¨ ⎧
1 x12 1[ ]
u1 = G 11 d x2 d x3 = J3 (x1 , c1 t) + J1 (x1 , c2 t) − x12 J3 (x1 , c2 t)
4 π c1 c2
R2
[ ( ( ))]}
−t I3 (x1 , c1 t) − I3 (x1 , c2 t) − 3x12 I5 (x1 , c1 t) − I5 x1 , c2 t ,
(9.92)

where:
¨ ¨
1 1
Jl (x1 , t) = δ(t − r )d x2 d x3 , Il (x1 , t) = H (t − r )d x2 d x3 . (9.93)
rl rl
R2 R2

/ (9.93) in polar coordinates ξ and α:


It is convenient to calculate integrals in
x2 = ξ cos α, x2 = ξ sin α, where ξ = x22 + x32 ≥ 0, −π < α ≤ π. At the
same time, we take into account that the/equation f (ξ) = t − r = 0 and inequality
f (ξ) ≥ 0, have solutions ξ = ξ0 = t 2 − x12 and ξ ≤ ξ0 only if t ≥ |x1 |. In
addition, according to the properties of the Dirac delta function:
|
1 r || t
δ(t − r ) = ' δ(ξ − ξ0 ) = | δ(ξ − ξ0 ) = δ(ξ − ξ0 ), r |ξ=ξ0 = t.
| f (ξ0 )| ξ ξ=ξ0 ξ0

Then, the integrals I 3 and I 5 in (9.93) are equal:

∫∞
t ξ
J1 (x1 , t) = H (t − |x1 |)2π δ(ξ − ξ0 )dξ = 2πH (t − |x1 |),
ξ0 r
0
∫∞
t ξ 2π
J3 (x1 , t) = H (t − |x1 |)2π δ(ξ − ξ0 )dξ = 2 H (t − |x1 |). (9.94)
ξ0 r 3 t
0

Integrals I 3 and I 5 are calculated taking into account the media of sub-integral
functions and replacing the integral variable y = ξ2 :

∫ξ0
ξ
I3 (x1 , t) = H (t − |x1 |)2π dξ = π H (t − |x1 |)
r3
0

∫ξ0
2

dy (t − |x1 |)
( ) = 2π H (t − |x1 |)
y + x12 3/ 2 t|x1 |
0
∫ξ0
ξ
I5 (x1 , t) = H (t − |x1 |)2π dξ = π H (t − |x1 |)
r5
0
276 9 Dynamic Problems of Solid Mechanics

∫ξ0
2
(3 )
dy t − |x1 |3
( )5 2 = 2π H (t − |x1 |) (9.95)
y + x12 / 3t 3 |x1 |3
0

Substituting (9.94) and (9.95) into (9.92), we finally get:


(. ) 1
u 1 x , t = u 1 (x1 , t) = H (c1 t − |x1 |).
2c1

Example 2 Find displacements in elastic space under the action of external forces
Fk = δ(t)δ(x1 , x2 )δkm , (k, m = 1, 2), F3 ≡ 0 and zero initial conditions.

In this case, the displacements are defined as:

∫∞
u k (→ x , t) ∗ ∗δ(t)δ(x1 , x2 )δlm =
x , t) = G kl (→ x , t)d x3 ,
G km (→
−∞
(9.96)
∫∞
u 3 (→ x , t) ∗ ∗δ(t)δ(x1 , x2 )δlm =
x , t) = G 3l (→ x , t)d x3 .
G 3m (→
−∞

Remark In this example, in (9.96) and beyond, all integer indices are 1 and 2.

The function G 3m is odd in coordinate x3 at m = 1, 2. Therefore, the corre-


sponding integrals in (9.96) are zeros and therefore u 3 ≡ 0 at this type of perturbation.
For the remaining components of the displacement vector, we get:

1 xk xm 1
uk = K 3 (x1 , x2 , c1 t) + [δkm K 1 (x1 , x2 , c2 t) − xk xm K 3 (x1 , x2 , c2 t)]
4π c1 c2
− t[δkm [L 3 (x1 , x2 , c1 t) − L 3 (x1 , x2 , c2 t)]
( ( ))]}
−3xk xm L 5 (x1 , x2 , c1 t) − L 5 x1 , x2 , c2 t (9.97)

where
∫∞ ∫∞
1 1
K l (x1 , x2 , t) = δ(t − r )d x3 , L l (x1 , x2 , t) = H (t − r )d x3 . (9.98)
rl rl
−∞ −∞

When calculating integral (9.98), we take into account that the equation t − r = 0
and inequality t −r ≥ 0 have solutions only /
when t ≥ r2 . These /solutions have these
forms: x3 = ±ζ0 and |x3 | ≤ ζ0 , where ζ0 = t 2 − r22 and r2 = x12 + x22 .
In addition, according to the properties of the Dirac delta function, we have:
|
r ||
δ(t − r ) = [δ(x3 + ζ0 ) + δ(x3 − ζ0 )]
|x3 | |x3 =±ζ0
9.11 Model Problems on Perturbation Propagation in Elastic Space 277

t
= [δ(x3 + ζ0 ) + δ(x3 − ζ0 )], r |x3 =±ζ0 = t
ζ0

Then for integrals K 1 and K 3 in (9.98), taking into account the even function of
the sub-integral function, we get:

∫∞
1
K 1 (x1 , x2 , t) = 2 δ(t − r )d x3
r
0
∫∞
t 1 ( )−1 2
= 2H (t − r2 ) δ(x3 − ζ0 )d x3 = 2 t 2 − r22 + /
ζ0 r
0
∫∞
1
K 3 (x1 , x2 , t) = 2 δ(t − r )d x3
r3
0
∫∞
t 1 2( )−1 2
= 2H (t − r2 ) δ(x3 − ζ0 )d x3 = 2 t 2 − r22 + / (9.99)
ζ0 r 3 t
0

The integrals L 3 and L 5 are calculated taking into account the even function of
sub-integral functions:

∫∞ ∫ζ0
1 d x3
L 3 (x1 , x2 , t) = 2 H (t − r )d x3 = 2H (t − r2 ) ( )3 2 =
r2 + x32 /
r 3 2
0 0
|ζ0
|
x3 | ( )
= 2H (t − r2 ) / | = 2 t 2 − r 2 1/ 2
| 2 2 +
r2 r2 + x2 |
2 2
t r2
3 0

∫∞ ∫ζ0
1 d x3
L 5 (x1 , x2 , t) = 2 H (t − r )d x3 = 2H (t − r2 ) ( )5 2
r5 r2 + x32 /
2
0 0
( ) |ζ0
x3 3r22 + x32 || r22 + 2t 2 ( 2 )1 2
= 2H (t − r2 ) ( ) | = 2 t − r22 +/ (9.100)
3r 4 r 2 + x 2 / |
3 2 3 4
3t r2
2 2 3 0

Substituting (9.99) and (9.100) into (9.97), we finally get:


⎧ [ ]
1 1( 2 2 )
2 1/ 2
( )
2 −1/ 2
u k = u k (x1 , x2 , t) = − δkm c t − r2 + − c2 t c2 t − r2 +
2 2 2
2π r22 c1 1

xk xm Σ2
(−1) j( )( ) ⎬
−1 2
+ 2 2c2j t 2 − r22 c2j t 2 − r22 + / . (9.101)
r2 cj ⎭
j=1
278 9 Dynamic Problems of Solid Mechanics

Example 3 Find displacements in elastic space under the action of the concentrated
force F applied at the origin of coordinates and directed along the axis Ox 1 and equal
to F = ρ f (t)H (t) and zero initial conditions.

Calculations are performed in dimensionless variables:

xi ui c∗ t ϑ Fi L ∂ ϑ'
xi' = ; u i' = ; τ = ; ϑ' = ; Fi' = 2 − .' ' ;
L L L T0 c1 ∂ xi
' σi j
σi j = (i, j = 1, 2, 3);
λ + 2μ
λ ν . T0 c∗
χ= = ; .' = ; γm = (m = 1, 2);
λ + 2μ 1−ν ρ c1
2 cm
/
c1 γ2 2
η= = = ;
c2 γ1 1−χ
/ / /
γ = γ1 ; r ' = r L; ϕ' = ϕ L 2 ; Φ '1 = Φ 1 L.
( / )
where u i , σi j , Fi , Fi − . ∂ϑ ∂ xi are, accordingly, components of displacement
vector,the stress tensor, the mass force vector, and the fictitious mass force vector; cm
are velocities of propagation of longitudinal and transverse waves in elastic medium;
λ, μ are Lamé constants; ν is the Poisson’s ratio; ρ is the medium density; . is the
coefficient of thermal expansion of the medium; T0 is the medium temperature in
initial condition; c∗ is a parameter having velocity dimension; L is the characteristic
linear dimension; ϕ and Φ 1 are respectively scalar potentials of displacement vector
and mass forces. /( )
We accept that: f (t) = f 0 = const; f 0' = f c12 L = 1; γ1 = 1; γ2 = 1, 871.
For the example under consideration, the coordinates of body force have the form:

F1 = f + (t)δ(→
x ), F2 = F3 ≡ 0; f + (t) = f (t)H (t).

The expression for the components of the displacement vector, taking into account
the conditions of the task, takes this form:

u k (→ x , t) ∗ ∗ f + (t)δ(x1 , x2 , x3 ) = G k1 (→
x , t) = G k1 (→ x , t) ∗ f + (t).

We substitute the influence function G k1 from (9.91) and we come to the following
equalities:
[ 2 ( ) ( ) ( )
1 x1 r 1 x12 r
u 1 (x1 , x2 , x3 , t) = f+ t − + 2 1 − 2 f+ t − +
4π c12 r 2 c1 c2 r c2
[ ( ) ( )]
x1 xk Σ
2
j 1 r 3 (2) r
u k (x1 , x2 , x3 , t) = − (−1) f+ t − + 2f t− ,
4π r 3 j=1 c2j cj r cj
9.12 Model Problems on Non-stationary Perturbation Propagation 279

where

∫t ∫t
(2)
f (t) = t+ ∗ f + (t) = (t − τ)H (t − τ) f (τ)dτ = H (t) (t − τ) f (τ)dτ.
0 0
(9.102)

The function f (2) (t) defined by formula (9.102) is a primitive function (the integral
of the function) f + (t).
From (9.102), for the function f (t), given by the conditions of the example, we
obtain:

∫t
(2) f0 2
f (t) = f 0 H (t) (t − τ) f (τ)dτ = t .
2 +
0

9.12 Model Problems on Non-stationary Perturbation


Propagation

In the case of problems for plane, taking into account the setting of problems
described in the previous section, we have:

u k = u k (x1 , x2 , t), u 3 ≡ 0; ε13 = ε23 = ε33 ≡ 0; σ13 = σ23 ≡ 0,


λ
σ33 = (σ11 + σ22 ).
2(λ + μ)

The equations of motion are written as:

( ) ∂θ
ü k = c12 − c22 + c22 δu k + Fk (x1 , x2 , t), x→ = (x1 , x2 ) ∈ R 2 , t > 0.
∂ xk
(9.103)

To construct a full resolution system of equations, we must add initial conditions


to Eq. (9.103):

u k |t=0 = ϕk (x1 , x2 ), u̇ k |t=0 = ψk (x1 , x2 ), (x1 , x2 ) ∈ R 2 .

Integral representations for displacements and stresses have the same form as
expressions (9.86) and (9.87).
The influence function G km (x1 , x2 , t) for two-dimensional problems was actually
constructed in Example 2 previously discussed (see expression (9.101)):
280 9 Dynamic Problems of Solid Mechanics
⎧ [ ]
1 1( 2 2 )
2 1/ 2
( )
2 −1/ 2
G km = − δ km c t − r +
− c2 t 2 2 2
c t − r +
2πr 2 c1 1 2

)−1 2 ⎬ / (9.104)
xk xm Σ (−1) j ( 2 2 )(
2
+ 2 2c j t − r 2 c2j t 2 − r 2 + / , r = x12 + x22 .
r cj ⎭
j=1

Example 4 Find displacements in elastic plane under the action of the concentrated
force F applied at the origin of coordinates and directed along the axis Ox 1 and equal
to: F = ρ f (t)H (t) and zero initial conditions.

Calculations are performed in dimensionless


/( ) variables (see example 3) and we
accept that: f (t) = f 0 = const; f 0' = f c12 L = 1; γ1 = 1; γ2 = 1, 871.
For the example under consideration, the body force coordinates have the form:

F1 = f + (t)δ(x1 , x2 ), F2 ≡ 0; f + (t) = f (t)H (t).

The expression for the components of the displacement vector, taking into account
the conditions of the example, takes the form:

u k (x1 , x2 , t) = G k1 (x1 , x2 , t) ∗ ∗ f + (t)δ(x1 , x2 ) = G k1 (x1 , x2 , t) ∗ f + (t). (9.105)

We obtain the influence function G k1 from (9.104):


⎧[ ]
1 1( 2 2 )1 2 ( )−1 2
G αα = − c1 t − r 2 +/ − c2 t 2 c22 t 2 − r 2 + /
2π r 2
c1

xα2 Σ (−1) j ( 2 2
2
)( ) ⎬
−1 / 2
+ 2 2c j t − r 2 c2j t 2 − r 2 + ;
r j=1 c j ⎭

x1 x2 Σ (−1) j ( 2 2 )( )−1 2
2
G 12 = − 4
2c j t − r 2 c2j t 2 − r 2 + /
2π r j=1 c j

We substitute these equations into (9.105) and we obtain that the displacements
are determined by the following formulas:
[ ( )
1 1 1 1
u 1 (x1 , x2 , t) = − N11 (r, t) − N (r, t) + N−1,2 (r, t)
2π c1 r 2 c2 r 2 12

( )
x 2 Σ (−1) j 2
2
+ 12 N 1j (r, t) + N −1, j (r, t) ⎦;
r j=1 c j r2
( )
x1 x2 Σ (−1) j 2
2
u 2 (x1 , x2 , t) = − N1 j (r, t) + N−1, j (r, t) ,
2π r 2 j=1 c j r2
9.13 Model Problems on Volume Perturbation Propagation 281

∫t ( )−1/ 2 ( ) ∫t f (t−τ)dτ
where N−1, j (r, t) = c2j τ2 − r 2 f (t − τ)dτ = H c j t − r √ 2 2 2,
+ / c j τ −r
0 r cj

∫t ( )−1/ 2 ∫t /
2 2 2 ( )
N1, j (r, t) = cjτ − r f (t − τ)dτ = H c j t − r f (t − τ) c2j τ2 − r 2 dτ.
+ /
0 r cj

We use a convolution operation to determine the displacements, then:

[
f0 Σ 1
2
x2 − x2 ( )1 / 2
u 1 (x1 , x2 , t) = f 0 G 11 (x1 , x2 , t) ∗ H (t) = (−1) j 2 4 1 t c2j t 2 − r 2
2π c 2r +
j=1 j
(( / )/ ) ( ]
1 )
+ ln c j t + c2j t 2 − r 2 r H cjt − r ,
2c j

f 0 x1 x2 Σ (−1) j ( 2 2 )
2
u 2 (x1 , x2 , t) = f 0 G 21 (x1 , x2 , t) ∗ H (t) = t c t − r 2 1/ 2 .
+ j
2π r 4 cj +
j=1
(9.106)

Remark Construct formulas (9.106) yourself.

9.13 Model Problems on Volume Perturbation Propagation

Consider the problem of spreading perturbations in a semi-space x1 ≥ 0, that are


uniformly distributed over the boundary plane x1 = 0. Half-space is an elastic
homogeneous isotropic medium.
The initial-boundary-value problem includes [Gorshkov, A.G. Waves in contin-
uous media: coursebook/A.G. Gorshkov, A.L. Medvedskiy, L.N. Rabinskiy, D.V.
Tarlakovskiy. – Moscow. Pub by “FIZMATLIT”, 2004. 472 p. (in Russian)]:
• Equation of elastic medium motion

γα2 ü α = u α,11 + ηα Fα (α = 1, 2, 3),

• Initial conditions

u α |τ=0 = ϕα (x1 ), u̇ α |τ=0 = ψα (x1 ),


282 9 Dynamic Problems of Solid Mechanics

• Boundary condition

( )|
α u + βu ,x |x=0 = q(τ).

Remark It can be shown that all types of solid mechanics boundary conditions can
be written in one general form:

( )| ( )|
α0 u i + β0 u i,x |x=0 = q0i (τ), α1 u̇ j + β1 u j,x |x=0 = q1 j (τ), (i, j = 1, 2, 3; i / = j )

Σ
1 Σ
1
where αk , βk ∈ R; k = 0, 1; α2k /= 0, β2k /= 0
l=0 l=0

Therefore, if α0 = 1, α1 = 0, β0 = 0, β1 = γ2j , and functions q0i (τ) = u 0i (τ),


q1 j (τ) = p0 j (τ), then there are mixed boundary conditions:
( )|
( u i )|x=0 = u 0i (τ), σ1 j |x=0 = p0 j (τ), (i, j = 1, 2, 3; i /= j).

Since the components of the displacement vector in this case are independent, it
is sufficient to construct a solution to the following problem:

γ2 ü = u ,11 + F,

u|τ=0 = ϕ(x), u̇|τ=0 = ψ(x),

( )|
α u + β u ,x |x=0 = 0, u(x, τ) = O(1), x → +∞.

Remark The components of the displacement vector are independent if they can
be found independently of each other and are not connected to each other through
boundary conditions.

We will use the definition of fundamental solutions. Then the displacements and
stresses for the problem under consideration can be represented as follows:

∫∞
u(x, τ) = G(x, τ; ξ) ∗ F(ξ, τ)dξ
0
⎡ ⎤
∫∞ ∫∞
+ γ2 ⎣ Ġ(x, τ; ξ)ϕ(ξ)dξ + G(x, τ; ξ)ψ(ξ)dξ⎦; (9.107)
0 0

∫∞
σ(x, τ) = G σ (x, τ; ξ) ∗ F(ξ, τ)dξ
0
9.13 Model Problems on Volume Perturbation Propagation 283
⎡∞ ⎤
∫ ∫∞
+ γ2 ⎣ Ġ σ (x, τ; ξ)ϕ(ξ)dξ + G σ (x, τ; ξ)ψ(ξ)dξ⎦,
0 0

where G(x, τ ; ξ) and G σ (x, τ ; ξ) are volumetric influence functions.


The first of these functions is the solution to the boundary-value problem:

γ2 G̈ = G ,x x + δ(τ)δ(x − ξ),

|
G|τ=0 = 0, Ġ |τ=0 = 0, (9.108)

( )|
α G + β G ,x |x=0 = 0, G = O(1), x → +∞.

The second function is written on the basis of the relation determining the stresses
through the displacements: /
G σ (x, τ; ξ) = G ,x (x, τ; ξ) ηα ..
We apply the Laplace time transform to the problem (9.108) and solve the
boundary-value problem with respect to images G L (x, s; ξ). As a result, we get
the following formulas for images of influence functions:

1
G L (x, s; ξ ) = {exp(γ s (x − ξ ))
2s γ

α + βγ s [ ]
− exp(−γ s (x + ξ )) − 2sh γ s (x − ξ ) H (x − ξ ) ;
α−βγ s
1
G σL (x, s; ξ ) = exp(γ s (x − ξ ))}
2ηα

α + βγ s [ ]
+ exp(−γ s (x + ξ )) − 2ch γ s (x − ξ ) H (x − ξ ) .
α−βγ s

The originals of these functions are easily defined using Laplace transform prop-
erties. It is more convenient to do this for each specific case of boundary conditions
on the surface x = 0.

Example 5 Determine the stress–strain state of the elastic half-space, in which at


the initial moment of time, the displacements equal to ϕ(x) = x e−x / 3 , and the
velocities equal to zero. The half-space boundary is fixed, there are no mass forces.
Put in γ = ηα = 1.

For this problem, the formulas for the stress–strain state component (9.107) have
the form:
∫∞
u(x, τ) = Ġ(x, τ; ξ)ϕ(ξ)dξ, (9.109)
0
284 9 Dynamic Problems of Solid Mechanics

∫∞
σ(x, τ) = Ġ σ (x, τ; ξ)φ(ξ)dξ. (9.110)
0

Time derived from influence functions is the differentiation of influence functions


of the following form:

1 { [ ] [ ]
G(x, τ ; ξ ) = H τ + γ (x − ξ ) − H τ − γ (x + ξ )

( [ ] [ ])}
−H (x − ξ ) H τ + γ (x − ξ ) − H τ − γ (x − ξ )
1 { [ ] [ ]
G α (x, τ ; ξ ) = δ τ + γ (x − ξ ) + δ τ − γ (x + ξ )
2 γ ηα
( [ ] [ ])
− H (x − ξ ) δ τ + γ (x − ξ ) + δ τ − γ (x − ξ )

Then
1{ [ ] [ ]
Ġ(x, τ ; ξ ) = δ τ + γ (x − ξ ) − δ τ − γ (x + ξ )
2 ( [ ] [ ])}
−H (x − ξ ) δ τ + γ (x − ξ ) − δ τ − γ (x − ξ )
1 { '[ ] [ ]
Ġ α (x, τ ; ξ ) = δ τ + γ (x − ξ ) + δ ' τ − γ (x + ξ )
2γ ηα
( [ ] [ ])}
−H (x − ξ ) δ ' τ + γ (x − ξ ) + δ ' τ − γ (x − ξ )

We substitute these functions into (9.109) and (9.110), and take into account
their media and initial conditions of the problem. As a result, we determine the
displacements and stresses:

1{ ( / )
u(x, τ ) = (τ + x) exp −(τ + x) 3 H (τ )
[2 ( / ) ( / )]}
+ (x − τ )+ exp −(x − τ ) 3 − (τ − x)+ exp −(τ − x) 3 ,
⎧( )
1 τ +x ( / )
σ (x, τ ) = 1− exp −(τ + x) 3 H (τ )
2 3
( )
τ −x ( / )
+ 1− exp −(τ − x) 3 H (τ − x)
3
( ) ⎧
x −τ ( / )
+ 1− exp −(x − τ ) 3 H (x − τ ) .
3

Control Questions
1. What is called a “wave in solid deformable media”?
2. Wave “carries” in solid:
1) energy;
2) material;
9.13 Model Problems on Volume Perturbation Propagation 285

3) both.
3. What is the difference between the concepts of “wave” and “oscillation”?
4. What are the main characteristics of the wave?
5. What are the definitions of longitudinal and transverse surface waves?
6. What are the definitions of plane waves and spherical waves?
7. What is the definition of Shock waves?
8. Wave equation:
1) ∂∂ xϕ2 = a 2 ∂∂tϕ2 ;
2 2

[ ]
2) εi j = 2G
1
σi j − 3ν
1+ν
σδi j ;
∂2ψ 2 ∂ψ 2∂ ψ
2
3) ∂r 2
+ r ∂r
= a ∂t 2 .
9. What is the complete system of resolving equations of the dynamic theory of
elasticity?
10. What is the Lamé equations of dynamic elasticity theory? / /

11. Propagation velocities of which waves are: c1 = (λ + 2μ)/ρ, c2 = μ ρ?
12. What is the boundary-value problem of dynamic theory of elasticity in case of
boundary conditions in stresses?
13. What is the Cauchy boundary problem of dynamic elasticity theory?
14. What are the Kinematic compatibility conditions of Lamé system?
15. What are the dynamic compatibility conditions of Lamé system?
16. What is the definition of Discontinuity surface?
17. What is the definition of Stokes problem?
18. D’Alembertian:
( ) ( ) ( )
1) ◻ a ∂ϕ = Δ ∂ϕ − a12 ∂t∂ 2 ∂ϕ
2

∂x ∂x ∂x
;
2) ϕ(t) = exp(i pt) = cos( pt) + i sin( pt);
3) σikj, j − ρωk2 u ik = 0.
19. What are the natural oscillations of elastic bodies? What is the fundamental
(natural) frequencies of elastic body?
20. What are the definitions of forced oscillations of elastic bodies?
21. What are the main features of plane wave propagation in unlimited elastic
bodies?
22. What are the main features of shock wave propagation in unlimited elastic
bodies?
23. What are the definitions of progressive waves?
24. What are the definitions of Rayleigh waves?
25. What are the definitions of Mach numbers?
Chapter 10
Mathematical Models of Special Classes
of Solid Mechanics Problems

One of the most important points in constructing mathematical models of a solid


deformable body state and behavior is the choice of constitutive equations between
components of stress–strain state (which determines the medium behavior: elasticity,
viscoelasticity, plasticity, etc.). Therefore, it is natural to wish to have an “Simulation
Software” that allows one to consider as wide a range of physical relationships as
possible to model the behavior of the real media.
Unfortunately, there are no universal, “life-friendly” methods and technologies.
Due to these circumstances, it is desirable to have several “application software”
implementing different laws of the behavior of solid media.
At the same time, another way is possible: the development of approaches that
allow modeling mechanical processes and phenomena to reduce real media to some
“effective media”. As an “effective medium” it is convenient to accept a model of
elastic medium, since it is obvious that to study the mechanical behavior of the
elastic medium, there is the largest number of well-studied and tested methods and
approaches.
Consider some of the approaches that allow you to perform this procedure.

10.1 Solving Problems for Heterogeneous Nonlinear Elastic


Media

10.1.1 Solving Problems for Physically Nonlinear Elastic


Medium

We consider the problems of the theory of elasticity, which are non-linear physic, but
linear geometry. In this case, the strains of extensions and shifts are small compared
to the unit.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 287
M. Zhuravkov et al., Mechanics of Solid Deformable Body,
https://doi.org/10.1007/978-981-19-8410-5_10
288 10 Mathematical Models of Special Classes of Solid Mechanics Problems

The relations between stresses σi j and small elastic strains ei j for a homogeneous
isotropic body according to a physically nonlinear theory can be written as follows:
( ) ( )
σ ii = 3K χ(e0 )e0 + 2μγ ψ20 (eii − e0 ) ; σ i j = μγ ψ20 ei j , (i /= j ) , (10.1)

where K = λ + 2μ/3 is the bulk modulus; e0 = (1/3)(e11 + e22 + e33 ) = eii /3 is


the average volumetric strain; ψ0 is shear strain intensity:
[( / )( 2 )
ψ20 = (4/3) 2 3 e11 + e22
2
+ e33
2
− e11 e22 − e11 e33 − e22 e33
( / )( 2 )]
+ 1 2 e12 + e13
2
+ e23
2
;
( )
where χ(e0 ) and γ ψ20 are respectively, extension and shift functions, which can be
represented by power series in the following form:

χ(e0 ) = 1 + χ1 e0 + χ2 e02 + χ3 e03 + · · · ,


( ) (10.2)
γ ψ20 = 1 + γ2 ψ20 + γ4 ψ40 + γ6 ψ60 + · · · .

Constants χi and γ2 j can be determined experimentally.

Remark For a large range of materials, the extension and shift functions with a
sufficient degree of accuracy can be expressed in the form:
( )
χ(e0 ) = 1 ; γ ψ20 = 1 + γ2 ψ20 .

We present the ratios (10.1) as:


(( / ) ( / ))
σ i j = λθδi j + μ ∂ u i ∂ x j + ∂ u j ∂ xi + 2μδηi j , (10.3)

where 0 ≤ δ ≤ 1; if δ = 0 then Eq. (10.3) moves to linear, and if δs = 1, move to


non-linear Eq. (10.1).
The perturbation function ηi j characterizes the deviation of physically nonlinear
relation (10.3) from the equations of the Hooke’s law. When the extension and shift
functions are represented in rows (10.2), perturbation function is as follows:
∞ [
Σ ( )]
1+ν
η ij = γ2n ψ2n
0 ei j + δi j χn e0n+1 − γ2n ψ2n
0 e0 . (10.4)
n=1
1 − 2ν

We substitute (10.3) into the Cauchy equilibrium equations. We obtain:

1 ∂θ Xi
Δu i + + + δFi = 0 , (i = 1, ..., m), (10.5)
1 − 2ν ∂ xi μ

Σ
m /
where Fi = 2 ∂ηi j ∂ x j .
j=1
10.1 Solving Problems for Heterogeneous Nonlinear Elastic Media 289

The procedure for solving problems according to the described approach is as


follows.
Let it be necessary to determine the stress–strain state of a physically non-linear
body. On one part of the boundary surface of the body Su , displacements u i0 are
specified, and on the rest part Sσ , external forces Ti0 are specified, S = Su ∪ Sσ :
|
Σ |
m
|
u i | Su = u i0 , σi j n j || = Ti0 . (10.6)
j=1 |

Represent u i , ei j , σi j , ψ20 , ηi j , Fi in rows by degree of parameter δ:

{ } ∞
Σ { }
u i , ei j , σi j , ψ20 , ηi j , Fi = δn u i(n) , ei(n)
j , σ (n)
ij , ψ (2n)
0 , η(n)
ij , Fi
(n)
. (10.7)
n=0

We add the corresponding series (10.7) to the Eq. (10.5) and equate the expressions
with the same degrees of the parameter δ. We get:

1 ∂θ(0) Xi
Δu i(0) + =− , (i = 1, ..., m); (10.8)
1 − 2ν ∂ xi μ

1 ∂θ(n)
Δu i(n) + = −Fi(n−1) , (n ≥ 1). (10.9)
1 − 2ν ∂ xi

Thus, in zero approximation we have a system of Lamé equations corresponding


to linear theory of elasticity. And in each subsequent approximation (n ≥ 1), the
right parts of the system (10.9) are known functions depending on the solution of the
problem in the previous approximations. They are defined by the formula:

Σ
m /
Fi(n−1) = 2 ∂ηi(n−1)
j ∂x j.
j=1

Boundary conditions (10.6) in the nth approximation will have the following form:
|
| |
| Σ
m
|
u i(n) | = u i0(n) ; σi j · n j || = Ti(n) . (10.10)
Su
j=1 |

To solve the problem (10.6), (10.8) in zero approximation and (10.9), (10.10) at
n ≥ 1, you can use some approaches to solve the problems of the theory of elasticity.
After the determination of u i(n) , strain components ei(n)
j are determined from
Cauchy ratios
290 10 Mathematical Models of Special Classes of Solid Mechanics Problems
( )
1 ∂u i(n) ∂u (n)
ei(n)
j
= + .
j
2 ∂x j ∂ xi

Stresses σi(n)
j are calculated by the formula:

σi(n) (n) (n) (n−1)


j = λ θ δi j + 2μ ei j + 2μ ηi j .

Thus, the physically non-linear boundary problem of the theory of elasticity was
reduced to a sequence of corresponding linear boundary problems.

10.1.2 Solution of Problems for Orthotropic Solids

We consider an elastic homogeneous solid, for which the connection between stresses
σi j and small strains ei j is described using the following relations:

σ11 = c11 e11 + c12 e22 + c13 e33 ;


σ22 = c21 e11 + c22 e22 + c23 e33 ;
σ33 = c13 e11 + c23 e22 + c33 e33 ;
σ12 = c66 e12 ; σ23 = c44 e23 ; σ13 = c55 e13 . (10.11)

We select a part in these ratios in the form corresponding to the Hooke’s law for
an isotropic solid: σi j = λθδi j + 2μei j . Then we present (10.11) in the following
form:

σ11 = (c11 − c12 )e11 + c12 (e11 + e22 + e33 ) + (c13 − c12 )e33 ;
σ22 = (c11 − c12 )e11 + c12 (e11 + e22 + e33 )
+ (c22 − c11 )e22 + (c23 − c12 )e33 ;
σ33 = (c11 − c12 )e11 + c12 (e11 + e22 + e33 ) + (c13 − c12 )e11
+ (c23 − c12 )e22 + (c33 − c11 )e33 ;
σ23 = μ e23 + (c44 − μ)e23 ; σ13 = μ e13 + (c55 − μ)e13 ;
σ12 = μ e12 + (c66 − μ)e12 . (10.12)

Acceptance for an orthotropic body is as follows:

μ = (1/2)(c11 − c12 ), λ = c12 . (10.13)

Then the ratios (10.12) can be written as:

σ i j = 2μei j + λθδi j + 2μδηi j (0 ≤ δ ≤ 1), (10.14)


10.1 Solving Problems for Heterogeneous Nonlinear Elastic Media 291

where perturbation functions ηi j characterize deviation of elastic properties of an


orthotropic solid from those of an isotropic solid. Functions ηi j according to (10.12)
have the form:
( / )
η11 = 1 2 (c13 − c12 )e33 ;
η22 = (1/2μ)[(c22 − c11 )e22 + (c23 − c12 )e33 ] ;
η33 = (1/2μ)[(c13 − c12 )e11 + (c23 − c12 )e22 + (c33 − c11 )e33 ] ;
η23 = (1/2μ)(c44 − μ)e23 ; η13 = (1/2μ)(c55 − μ)e13 ; η12 = (1/2μ)(c66 − μ)e12 .
(10.15)

At δ = 0, the ratio (10.14) turns into the Hooke’s law equations for an isotropic
medium, and at the δ = 1, in the Eq. (10.11) of the generalized Hooke’s law for an
orthotropic medium. Characteristics μ and λ are determined according to (10.13).
Expressions (10.13) can also be written using “technical constants”:

μ = (E 1 /2)[1 − ν12 − ν23 (ν13 + ν23 )(E 2 /E 3 )]·


[ ( 2 ) ]−1
1 − ν12
2
(E 1 /E 2 ) − ν23
2
(E 2 /E 3 ) − ν13 + 2ν12 ν13 ν23 (E 1 /E 3 ) ;
[ ]
2 −1
λ = (ν21 + ν31 ν23 )E 2 1 − 2ν12 ν23 ν13 − ν13 2
− ν122
− ν23 .

Let’s describe the task-solving procedure based on this approach.


Let it be necessary to determine the stress–strain state of the orthotropic solid
under given boundary conditions. We represent unknown components of the stress
tensor σi j and displacement vector u i in rows:

{ ∞
} Σ { }
σi j , u i = δk u i(k) , σi(k)
j . (10.16)
k=0

Note that the type of Eqs. (10.3) and (10.14) is identical. In addition, the selection
of unknowns u i and σi j in form (10.16) and (10.7) is also identical. Therefore, the
method of finding the components u i(k) and σi(k)
j in this case is similar to the method
described in the previous Sect. 10.1.1. The difference is that instead of the ratio
(10.4), perturbations ηi j should be determined on the basis of the formulas (10.15).

Remark This method of perturbing isotropic elastic properties allows you to find
solutions with the highest degree of accuracy for orthotropic bodies with elastic
properties that deviate slightly from the properties of the isotropic medium.
292 10 Mathematical Models of Special Classes of Solid Mechanics Problems

10.2 Solving the Problems of Linear Theory


of Viscoelasticity for Inhomogeneous Solids

This section describes a possible approach to account for rheological effects for inho-
mogeneous solids. We show that the problems of the linear theory of viscoelasticity
for inhomogeneous bodies can be reduced to solve the problems of the theory of
elasticity of inhomogeneous bodies.
The relationship between stresses and strains for an anisotropic inhomogeneous
linear viscoelastic body is determined by the relationships in the following form:

∫ t
σi j = λi jkl (t − τ, xs )dεkl (τ). (10.17)
0

Applying to (10.17), Laplace transform is as follows:

∫ ∞
f ( p) = f (t) exp(− pt)dt. (10.18)
0

As a result, we have:

σi j = C i jkl (xs )εkl , C i jkl = pλijkl (xs ).

In this case, for the displacement transformants u k , the equilibrium equation in


the differential form and the boundary conditions are written as:
( )
∂ ∂u k
C i jkl + ρX i = 0; (10.19)
∂x j ∂ xl
∂ u k || |
C i jkl (xs ) n j Sσ = q i (xs ) ; u i | Su = ϕi (xs ). (10.20)
∂ xl

Note that the relations (10.19) and (10.20) are valid if the boundary of the body S
and the boundary between the parts Sσ and Su , the surface S does not change over
time.
Thus, in these assumptions, the problem of deformation of a viscoelastic inhomo-
geneous solid is reduced for Laplace-transformed values to the boundary problems
(10.19) and (10.20) of the theory of elasticity of inhomogeneous solid.
Consider, for example, the problem of deforming an isotropic inhomogeneous
viscoelastic body. Let the volumetric properties of the material be perfectly elastic,
and the shear properties are viscoelastic. Then:
10.2 Solving the Problems of Linear Theory of Viscoelasticity … 293

∫ t
si j = 2Gei j − (t − τ)ei j (τ)dτ; θ = σ/K , (10.21)
0

where si j , ei j are stress and strain tensor deviators, i.e.,

si j = σi j − σ σi j ; ei j = εi j − θδi j /3. (10.22)

From (10.21) and (10.22), the stress tensor can be obtained:

∫ t
( ) ( )
σi j = K θδi j + 2G εi j − θδi j /3 − (t − τ) εi j − θδi j /3 dτ. (10.23)
0

Apply to (10.23) Laplace transform (10.18), and determine the values σi j =


λθδi j + 2μεi j , where λ = K − (2/3)G + (1/3)∂( p); μ = G − (1/2)∂( p) .
Therefore, a boundary-value problem for determining displacement transformants
using Laplace-transformed equations of equilibrium and boundary conditions is
brought to such boundary-value problem of elasticity theory of inhomogeneous
media relative to displacement transformants.
( )
( ) ∂θ ∂λ ∂μ ∂u i ∂u j
λ+μ + μΔu i + θ + + + ρX i = 0;
∂ xi ∂ xi ∂ xi ∂ x j ∂ xi
[ ( )]
∂ ui ∂ uj | |
λθδi j + μ + n j | Sσ = q i (xs ); u i | Su = ϕi (xs ).
∂ xj ∂ xi

In turn, we give one of the possible options for constructing resolving systems of
equations of problems of the theory of elasticity of inhomogeneous media.
The boundary-value problem of the theory of elasticity for an inhomogeneous
body is formulated as follows. You want to define a vector u(x) that satisfies in the
volume D = D ∪ S, the equilibrium equation is in the following form:
( )
∂θ ∂μ ∂u i ∂u j ∂λ ( )
(λ + μ) + μΔu i + + +θ + ρX i = 0, i = 1, m
∂ xi ∂x j ∂x j ∂ xi ∂ xi
(10.24)

under such boundary conditions at the S:


[ ( )]
∂u k ∂u i ∂u j | | ( )
λ δi j + μ + n j | Sσ = qi (xs ); u i | Su = ϕi (xs ), i = 1, m ,
∂ xk ∂x j ∂ xi

where λ = λ(xs ) and μ = μ(xs ) are Lamé constants; n j are components of the outer
normal vector restored at the point xs of the surface Sσ ; Sσ and Su are parts of surface
294 10 Mathematical Models of Special Classes of Solid Mechanics Problems

S with specified external loads qi and displacements ϕi , respectively; S = Sσ + Su ;


m = 2 or m = 3 is space dimension.
Differential equilibrium Eq. (10.24) can be written as:
( )
1 ∂θ Xi θ ∂λ 1 ∂μ ∂u i ∂u j
Δu i + = −ρ − − + .
1 − 2ν(x) ∂ xi μ(x) μ(x) ∂ xi μ(x) ∂ x j ∂x j ∂ xi
(10.25)

The representation of Eq. (10.25) corresponds to the Lamé equilibrium equations


of the theory of elasticity.
It should be noted that in “pseudo-elastic” problems, the vector u(x ) means
either the displacement vector transformant u(x ) or the vector of elastic fictitious
displacements u y (x ) . Accordingly, in this case, the components of either the stress
tensor transformant σi j or the stress tensor itself σi j are determined based on the
Hooke’s law.

10.3 The Theory of Creep of Isotropic Hardening Media

Remark The material in this section is based on work [Golub, V.P. To the creep
theory of isotropic initially hardening media / V.P. Golub // Applied mechanics,
1989. V.25. N2. P.90–100. (in Russian)].

The properties of the real medium in the constitutive creep equations are usually
given by one-dimensional models of various structures. Therefore, consider the
solution of the problem for the one-dimensional case when stresses and strains
are not tensors, but scalars.
The initial state of the isotropic medium with simple loading for one-dimensional
problems can be represented by the following equation:

σ = ϕ0 (ε) = g(ε)ε. (10.26)

We consider hardening media that are stable in the sense of M. Drucker’s postulate
and satisfy such condition:

dσ dε p > 0, (10.27)

where σ is the true tensile stress value; ε is the strain corresponding to this stress;
ϕ0 (ε) is the strain function, generally non-linear; g(ε) is the plastic modulus; dε p is
the increment of the plastic component of strain. For the function g(ε), we can write
it in the following expression:
⎧ ( )
E 0 ≤ ε ≤ εy ;
g(ε) = / ( ) (10.28)
ϕ0 (ε) ε y < ε ≤ εb ,
10.3 The Theory of Creep of Isotropic Hardening Media 295

Fig. 10.1 Creep curves t=0

C
B t=constant

A D

/
where ϕ0 (ε) > 0 is the derivative from ϕ0 (ε) to ε; E is the modulus of elasticity;
ε y , εb are strain values corresponding to yield strength σ y and temporary resistance
σb .
To build a creep model, the principle of similarity of isochronous creep diagrams,
formulated by Yu. N. Rabotnov, is used. Isochronous creep diagrams where mean
creep diagrams rebuilt in coordinates “σ − ε” by time parameters (Fig. 10.1).
The whole collection of isochrones is quite well approximated by the expression
in the following form:
( )−1
σ = ϕ(ε)θ(t) = ϕ(ε) 1 + a t b , (10.29)

where a, b are factors determined experimentally; ε = ε0 + εc .


In the general case, the explicit form of the function θ (t) is given by the weight
function of the linear-hereditary equation so that at t = 0 the function (10.29)
describes the elastic part ϕ0 (ε) of the instantaneous deformation diagram. Accord-
ingly, the similarity in the frame (10.29) is defined as the possibility of shifting from
diagram ϕ0 (ε) to isochrones along the abscissa ε at a given stress σ0 (Fig. 10.2,
dash-and-dot curve).
Equation (10.29) is essentially a relaxation equation.
Another formula of the similarity condition of isochronous creep diagrams may
be as follows. With a fixed value ε, the instantaneous values of the functions of the
scleronomic ϕ0 (εs ) and rheonomic ϕ0 (εr ) components of the complete strain are
equivalent and the following condition is met:
( )| ( ) ( )
ϕt εr |ε=const = ϕ0 εs = σ 1 + K ∗ , (10.30)
/
where εr = ε0 + εc ; ε0 = σ E is the elastic strain; εs = ε0 + εe if εs < ε y and
εs = ε y + ε p if εs > ε y ; K ∗ is the integral operator.
296 10 Mathematical Models of Special Classes of Solid Mechanics Problems

Fig. 10.2 Principle of t=0


similarity of creep curves C
A
t1
А
t2
A
D
A

From the condition of similarity (10.30), after simple transformations, taking into
account (10.26), a creep equation is obtained, relative to εc :
( )
ϕt (εr ) − ϕ0 ε y ( )
ε = c
+ ε y − ε0 . (10.31)
g(εs )

In Eq. (10.31), the creep strain is given by the following relation:


⎧ ( )
εe − ε0 ε0 ≤(εe ≤ ε y ; )
εc = (10.32)
ε y + ε p − ε0 0 ≤ ε p ≤ εb − ε y .

Determining the full creep strain rate as a partial time derivative of the expression
(10.31) leads to the following equation:
[ ( )]
∂ ϕt (εr ) − ϕ0 ε y 1
ε̇ =c
. (10.33)
∂t g(εs )

The Eq. (10.33) in the general case of variable loading, taking into account the
history of loading and the ratio (10.28), is converted to the type:
[ ( )] [ ]
∂ ψ(σ(τ), (t − τ)) − ϕ0 ε y ∂ ϕ0 (εs )
ε̇ =
c
, (10.34)
∂t ∂ εs

where ψ is the creep function, defined by relation (10.30).


The practical use of the governing Eqs. (10.33) and (10.34) assumes the specificity
of weight functions and function ϕ0 (s) . This corresponds to the choice of the core
structure of the integral operator (the selection should satisfy the best agreement
with the creep experiment). In the simplest case, the integral operator core can be
10.3 The Theory of Creep of Isotropic Hardening Media 297

selected in the form (10.29). The structure of the function ϕ0 (εs ) is determined from
the results of processing experimental diagrams of tension curves.
The similarity condition (10.30) is fundamentally different from (10.29), since
the equivalence of the isochronous diagram and the diagram of instantaneous strain
is not given by abscissa ε, but by ordinate ϕ(ε) and linear displacement is replaced
by a non-linear function (Fig. 10.2, dash-and-dot curve).
For real hardening media, differentiate (10.26) by ε and consider (10.27). As a
result, we can write:
/
dσ dε − σ/ ε = g ' (ε)ε ≤ 0.
/ /
Therefore, dσ dε > 0 or dσ dε = const.
As a result, the general case of hardening can be reduced to linear and non-linear
laws. The non-linear law is most often given by a power function.
Let’s consider features of medium creep with various nature of hardening defor-
mation. As a creep function, we choose the simplest version of the kernel in the form
similar to (10.29). For the creep rate ε̇c at constant stress σ from (10.33), we obtain:

dεc σ 1
=ab , (10.35)
dt g(ε) t 1−b

where a > 0; 0 < b < 1.


For the isotropic medium with linear hardening in the elastic region g(εs ) = E,
and plastic region g(εs ) = E ∗ = const, and E ∗ < E. In this case, the creep rate
according to (10.35), in the whole variation range t, is a decreasing function of time
(at t → ∞, ε̇c → 0) and the creep curve has no inflection because the second
derivative equals to

d 2 εc σ 1
= a b(b − 1) s 2−b
dt 2 g(ε ) t

and always less than zero (b − 1 < 0).


Therefore, the creep of the linearly hardening medium is unsteady (Fig. 10.3,
curve 1).
In solids with a clearly defined transition from the elastic section to the linear
hardening section, the effect of a hopping increase in creep rate occurs at a constant
stress (Fig. 10.3, curve 2). The effect is caused by a hopping change in the plasticity
modulus g(ε) at the moment of this transition to the value Δ g(ε) = E − E ∗ and it is
realized when the creep strain εc reaches the value εT − ε0 . It can also be seen from
(10.35) that the smaller the hardening modulus E ∗ , the higher the creep rate at this
point and at E ∗ → 0 (non-hardening media) ε̇c → ∞ (curve 3).
In the case where there is a transition region in the instantaneous deformation
diagram ϕ0 (εs ) between the elastic section and the hardening section, there is no
effect of the hopping creep rate (curve 4).
298 10 Mathematical Models of Special Classes of Solid Mechanics Problems

3
2

0 t

Fig. 10.3 Creep curves of linearly hardening medium

For an isotropic medium with power hardening, the creep determined by the value
g(εs ) = E ∗ = const in the region, where εc < ε y − ε0 , as in the previous case, is
unsteady (Fig. 10.4, curve 1). In the area of strains where εc > ε y − ε0 , for creep rate
from (10.33), we obtain:
( ) m1
dεc 1 σ[ ( ) ] 1−m 1
= ab σ 1 + at b − EεT m 1−b . (10.36)
dt B m t

For the second derivative:


( ) m1
d 2 εc 1 σ[ ( ) ] 1−m
= ab σ 1 + at b
− Eε T
m

dt 2 B m
⎧ ⎫
(1 − m) abσ t b 1
( ) − (1 − b) 2−b .
m σ 1 + at b − EεT t

2
3

Х
Х

0 t

Fig. 10.4 Creep curves of isotropic medium with power hardening


10.3 The Theory of Creep of Isotropic Hardening Media 299

This second derivative at a time t ∗ equals to


[ ]1
∗ (1 − b) ( / ) b
t = ( / ) · 1 − σy σ , (10.37)
a b m−1

and takes a zero value.


In (10.36)–(10.37) B > 0, m < 1 are parameters of hardening in power law.
Therefore, there is an inflection point on the creep curve, the creep rate after
the inflection point becomes increasing and the system of Eqs. (10.35) and (10.36)
describe all three creep stages (Fig. 10.4, curve 2).
For non-linear elastic media (σ y = 0)
[ ] b1
∗ (1 − b)
t = ( / )
a b m−1

and inflection points on the creep curve are independent for stress.
An additional condition for the existence of inflection points on the creep curve
according to (10.37) is the realization of inequality b > m, since σ < σ y (the initial
state is elastic). With b < m, Eq. (10.37) loses meaning and the inflection point is
determined by the moment of the creep strain εc reaches a value ε y − ε0 (curve 3).
Initial deformation hardening of medium is taken into account by introduction of
/ /
hardening speed ϕ0 (εs ) into governing Eqs. (10.33) and (10.34). As the value ϕ0 (εs )
decreases, the creep speed ε̇ increases (ε̇ → ∞). Then the condition:
c c

|
∂ ϕ0 (εs ) ∂ ϕt (εr ) ||
= →0 (10.38)
∂ εs ∂ εr |t=const

can be considered as a criterion for creep failure. The condition (10.38) corresponds
to the moment of the loss of deformation stability and is implemented in medium
with power hardening.

10.3.1 Consider the Situation of a Complex Stress State

The solution of creep problems in a complex stress state within the framework
of the approaches considered in this paragraph is based on the hypothesis of a
“uniform creep curve”. An obvious condition for the validity of such and similar
hypotheses is the implementation of the principle of similarity. When the theory is
being constructing, it is effective to use the similarity of isochronous creep diagrams
and diagrams of instantaneous strains, which makes it possible to generalize a one-
dimensional creep model (10.33) to a complex stress state based on the hypothesis
of “uniform deformation curve”.
300 10 Mathematical Models of Special Classes of Solid Mechanics Problems

The hardening medium Eq. (10.26) in this case is written as:

σ = ϕ(ε) = g(ε)ε, (10.39)

where σ , ∈ are the generalized stress and strain; ϕ(ε) is the known function of
material properties; g(ε) is the function independent of the type of stress state and
g(ε) ≡ g(ε) .

Remark Stress and strain intensities can be used as σ and ∈, respectively, as well as
the octahedral tangent stress and octahedral shift.

When constructing the constitutive creep equations for three-dimensional stress


and strain fields, we use the following assumptions:
• Medium is considered incompressible up to destruction;
• The law of plastic flow (10.27) is assumed to be fair, but does not contain hidden
variables;
• The generalized creep rate is associated with the generalized stress of the same
relationship as in the one-dimensional case.
Thus, for the creep rate ε̇icj in the general case of the stress state, based on (10.27),
(10.33), and (10.39), we obtain:
( )
3 ε̇ic (σi ) 1
ε̇icj = − σi j − δi j σi j ,
2 σi 3
[ ( r) ]
∂ ϕ ε
t i − σi 1
ε̇ic = , (10.40)
∂t g(εi )

where ε̇ic is the intensity of strain rates; σi is the stress intensity; δi j is the Kronecker
delta. In (10.40), the Huber-Mises plasticity condition was used.
Therefore, the procedure for constructing the theory of creep is described, based
on the refined principle of similarity of isochronous diagrams. This theory can be seen
as an attempt to mechanically generalize the concepts of equations of state and non-
linear heredity. The theory allows you to take into account the initial deformation
hardening of the medium, describe the third stage of creep, take into account the
history and cycle of loading.

10.4 Bilinear Theory of Elasticity

Determining constitutive equations of bilinear theory of elasticity in case of complex


stress state takes into account plastic compressibility of material so that relations
between strains and stresses for plastic bodies are bilinear [Nowacki, W.K. Stress
waves in non-elastic Solids \ W.K. Nowacki. – Pergamon Press Ltd, 1978. 247 p.].
10.5 Nonlinear Viscoelastic Media 301
/
These ratios for the active plastic loading process (when d J2 dt > 0, where J2 =
0.5 si j si j is the second invariant of the stress tensor deviator) have the form:
( )
1 1 1
ei j = si j + − (si j − si0j ),
2μ1 2μ2 2μ1
( )
1 1 1
εi j = σi j + − (σi j − σi0j ),
3K 1 2K 2 2K 1

where σi0j is the initial stress tensor corresponding to material transition points in
plastic state; μ1 ,K 1 are shape and volumetric deformation modulus respectively;
μ2 ,K 2 are plastic material constants:
/ /
μ2 = E 2 (2(1 + ν2 )), 3K 2 = E 2 (1 − 2ν2 ). (10.41)

As it is known, the coefficient ν2 is determined from the experience of uniaxial


p/ p
tension as the ratio of transverse plastic strain to longitudinal strain:ν2 = −ε2 ε1 .
Thus, ν2 and μ2 , respectively, in the ratios (10.41), play the role of Poisson’s ratio
and shear modulus in the plastic region.
The equations of bilinear theory in the case of a uniaxial stress state pass to the
relations of the deformation theory of plasticity.
The application of bilinear theory in complex stress state problems has the advan-
tage with respect to other plasticity theories that its equations are equally integrated
in both elastic and plastic regions (due to the same linear relationships between strain
and stress deviators and spherical components of tensors in both elastic and plastic
deformation regions). This is the theoretical convenience.
At the same time, this theory is associated with some simplifications of the physical
nature of problems.

10.5 Nonlinear Viscoelastic Media

Currently, the mathematical apparatus of linear viscoelasticity has been developed


quite fully. Therefore, linear viscoelastic models are widely used in practical appli-
cations. Attempting to describe the deformation behavior of bodies with non-linear
viscoelasticity led to the need to generalize the principle of Boltzmann superposition
to the superposition of responses, taking into account the history of deformation.
Thus, in a number of works for materials having similar stress relaxation curves, it
is proposed to use an equation in the following form [Ferry, John D. Viscoelastic
properties of polymers / John D. Ferry. – New York–London, 1961. 535 p.]:

∫ t
σ(t) = ϕ(ε(t)) − R(t − Θ)ϕ(ε(Θ))dΘ
0
302 10 Mathematical Models of Special Classes of Solid Mechanics Problems

where ϕ(ε(t)) is the empirical strain function.


For materials that have similar isochronous curves (i.e., dependencies built from
relaxation curves at fixed t), the hereditary non-linear theory of viscoelasticity, based
on the Rabotnov’s equation, has become very widespread. General view of the
Rabotnov’s equation:

∫ t
ψ[σ(t)] = ε(t) − R(t − Θ)ε(Θ)dΘ,
0

where ψ[σ] is the empirical stress function.


In general, the non-linear theory of viscoelasticity can be represented by the
Fréchet-Volterre’s equations, a special case of which is the non-linear principle of
superposition developed by Persaud, who generalized Boltzmann’s postulates:

∫ t
σ(t) = Eε(t)− R(t − Θ, ε(Θ))dΘ,
0

where R(t, ε) is the non-linear relaxation core.


Another equation of the non-linear theory of viscoelasticity is the Moskvitin’s
equation. This equation takes into account the accumulation of damages using a
linear Voltaire’s equation with time modified from the accumulated damages:

∫ t
ϕ(εu )K = f (σu , σ)σ + U (t − τ) f (σu , σ)σ(τ)dτ,
0

where ϕ(εu ) is the required function from strains.


The presence of nonlinearity in viscoelasticity equations as a function of strain
or stress significantly complicates the problem compared to the linear version of
viscoelasticity equations. Consider one of the most common and rather simple
approaches to considering the nonlinear viscoelastic properties of materials.
Constitutive equations between stresses and strains are taken as:

∫ t
si j = 2G( f (εu ) C- i j − R(t − τ) f (εu )i j (τ)dτ), σ = 3K ε, (10.42)
0

where si j C- i j , are deviators of stress and strain tensors; σ, ε are respectively, the
average stress and strain; G, K are instantaneous elastic moduli of shear and
volumetric strain; R(t) is the relaxation function.
Physical non-linearity of a material describes a universal function f (εu ), wherein
f (εu ) = 1, if the strain intensity εu does not exceed a certain threshold value εs :
εu ≤ εs .
10.5 Nonlinear Viscoelastic Media 303

Fig. 10.5 Results of creep


tests at pure shear

Physical Eq. (10.42) express the following: after the strain intensity at the point
reaches a certain threshold value εs , the strain field ∋ij at a given moment of time
is determined not only by instantaneous and preceding stress values sij , but also
depends on the type of the deformed state itself. Volume change is assumed to be
elastic. By resolving (10.42) with respect to sij , we obtain:

∫ t
2G f (εu ) C- i j = si j + Γ(t − τ )si j (τ)dτ, 3K ε = σ, (10.43)
0

where Γ(t) is the creep kernel.


Consider the technique for determining the non-linear function f (εu ) from exper-
imental creep curves. Let, for example, the results of creep experiments at pure shear
be known (Fig. 10.5).
In this case, six Eqs. (10.43) will be reduced to one, since ∋ij = 0 for all values
i and j, except for one component ∋12 = ε12 , similar to sij = 0 except for s12 = σ12 .
Where( 2/ ε12 is)the relative shear; σ12 is the corresponding shearing stress. In our case,

εu = 2 3 ε12 .
The lower curve in Fig. 10.5 corresponds to the linear behavior of the material
(0)
(εu ≤ εs ) at constant stress σ12 . The ratio is true along it:
⎛ ⎞
∫ t
2Gε(0)
12 (t) = (0) ⎝
σ12 1 + Γ(t − τ)dτ⎠. (10.44)
0

This curve can also be used to determine the creep kernel:

∫ t
2Gε(0)
12 (t)
Γ(t − τ)dτ = (0)
− 1.
σ12
0

For other curves from (10.43) follow:


304 10 Mathematical Models of Special Classes of Solid Mechanics Problems
⎛ ⎞
∫ t
(k) (k) ⎝
2G f (ε(k)
u )ε12 (t) = σ12 1 + Γ(t − τ)dτ⎠, (10.45)
0

(k)
where σ12 is constant stress along the k th curve (k = 1, 2, 3).
The left and right parts of Eq. (10.45) are divided into the corresponding parts of
the relation (10.44). After elementary transformations, we get experimental values
of the non-linear function at different points in time:
/( )
(k) (0) (0) (k)
f (t) = σ12 ε12 (t) σ12 ε12 (t) .

Since the strain intensity εu (t) and the non-linear function f (t) are known at each
time, it is possible to construct an experimental curve f ∼ εu by comparing their
values for the same t. Then constants are determined in accepted approximate formula
for non-linear function.
Note that the non-linear constitutive equations between stresses and strains can
be taken in another form:

∫ t
2G C- i j = g(σu )si j + Γ(t − τ)g(σu )si j (τ)dτ, 3K ε = σ, (10.46)
0

where g(σu ) is the universal stress intensity function determined experimentally.


Equations (10.42) and (10.46) are not equivalent to each other. One or another
method of accounting for the material physical non-linearity is selected depending
on the properties of the material and the available experimental data.

10.6 Method of Successive Approximations in Viscoplastic


Problems

Many materials, especially at high temperatures, show clearly rheonomic properties


in addition to plasticity. Such materials are called viscoplasticity.
Let’s consider one model that allows you to describe this behavior of mate-
rials. Constitutive equations in the presence of temperature field T (x,t) assume the
following:
⎛ ⎞
∫ t
si j = 2G(T )⎝ϕ(εu , T ) C- i j − R(t − τ) f (εu , T ) C- i j (τ)dτ⎠, σ = 3K (T )(ε − αT ),
0
(10.47)
10.6 Method of Successive Approximations in Viscoplastic Problems 305

where ϕ(εu , T ) is the plastic function; f (εu , T ) is the universal function of


rheonomic non-linearity of materials.
Inverse to (10.47), they are expressions of strains through stresses:

∫ t
2G(T ) C- i j = ϕ1 (σu ,T )si j + Γ(t − τ, T )g(σu , T )si j (τ)dτ,
(10.48)
0
3K (T )(ε − αT ) = σ.

In (10.48), it is noted that instantaneous modulus of elasticity G(T ) and


K(T ), creep kernel Γ(t,T ), and relaxation kernel R(t,T ), as well as plastic
functions ϕ(εu , T ), ϕ1 (σu , T ) and universal functions of physical non-linearity
f (εu , T ), g(σu , T ) can depend on temperature. At the same time ϕ(εu , T ) = 1
, if εu ≤ ε y (T ); ϕ1 (σu , T ) = 1 , if σu ≤ σ y (T ); f (εu , T ) = 1, if εu ≤ εs (T );
g(σu , T ) = 1, if σu ≤ σs (T ).
For a general statement of the problem, it is necessary to attach to equilibrium
Eqs. (10.47) or (10.48) (which are valid only with active loading), Cauchy relations,
and boundary conditions:
⎛ ⎞
∫ t
si j = 2G(T )⎝ϕ(εu , T ) C- i j − R(t − τ) f (εu , T ) C- i j (τ)dτ⎠, σ = 3K (T )(ε − αT ),
0
/
σi j , j +ρFi = 0, εi j = (u i , j +u j ,i ) 2, (10.49)

u i = u i0 (x) on Su , σi j l j = Ri on Sσ .
The solution of viscoplastic problems is associated with the solution of a system of
non-linear integra-differential partial differential Eqs. (10.49), which is an extremely
complex mathematical problem. Therefore, it is almost impossible to build an analyt-
ical solution to this problem. Therefore, you can use the method of successive approx-
imations (allocation procedure), which is based on the method of Ilyushin elastic
solutions.
We present the functions of plastic and physical non-linearity in the following
form:

ϕ(εu , T ) = 1 − ω(εu , T ), f (εu , T ) = 1 − ω1 (εu , T ), (10.50)

where 0 ≤ ω < 1, 0 ≤ ω1 < 1; and ω(εu ) = 0, if εu ≤ ε y .


Substitute (10.50) into (10.49). The physical relationships then become the
following:
306 10 Mathematical Models of Special Classes of Solid Mechanics Problems
⎛ ⎞
∫ t
si j = 2G(T )⎝(1 − ω(εu , T )) C- i j − R(t − τ)(1 − ω1 (εu , T )) C- i j (τ)dτ⎠,
0
σ = 3K (T )(ε − αT ).
(10.51)

Thus, with ω = ω1 = 0, Eqs. (10.51) describe linear viscoelastic properties of


materials.
We write the differential equilibrium equations and boundary conditions by
decomposing the stress tensor into the deviatoric and spherical parts:

si j , j +σ,i +ρFi = 0, (10.52)

u i = u i0 (x) on the Su , si j l j + σli = Ri on the Sσ (10.53)

The problem of viscoplasticity is solved in displacements. We substitute the


components of the deviatoric and spherical parts of the stress tensor (10.51) into
the equilibrium Eqs. (10.52) and force boundary conditions (10.53), and take into
account the Cauchy relations. As a result, we get the following generalized Lamé
equations and boundary conditions:

(λ + μ)θ,i +μΔu i + ρFi − Fω i − 3K αT,i = 0,

λθli + μ(u i , j +u j ,i )l j = Ri + Rωi + 3K αT li ,

where
⎛ ⎞
∫ t
Fω i = 2G ⎝ω C- i j + R(t − τ, T ) f (εu , T ) C- i j (τ)dτ⎠, j ;
0
⎛ ⎞
∫ t
Rω i = 2G ⎝ω C- i j + R(t − τ, T ) f (εu , T ) C- i j (τ)dτ⎠l j .
0

We take zero approximation as ω(0) = f (0) = 0. Then F ωi = Rωi = 0 and


to determine the first approximation ui (1) , we have the usual problem of linear
thermoelasticity. Other values are determined by displacements ui (1) :

C- i(1) (1)
j , εu , ω
(1)
≡ ω(ε(1)
u , T ), f
(1)
≡ f (ε(1) (1) (1)
u , T ), Fωi , Rω i .

For any kth approximation, the following equilibrium equations and boundary
conditions occur:
10.6 Method of Successive Approximations in Viscoplastic Problems 307

(λ + μ)θ,i(k) +μΔu i(k) + ρFi − Fωi


(k−1)
− 3K αT,i = 0, (10.54)

λθ(k)li + μ(u i(k) , j +u (k) (k−1)


j ,i )l j = Ri + Rωi + 3K αT li , (10.55)

where Fω(k−1)i , Rω(k−1)


i are known since they are determined by the previous
approximation (k − 1).
Equation (10.54) and boundary condition (10.55) are linear with respect to
unknown displacements ui (k) . They differ from the corresponding equations of the
theory of thermoelasticity in those fictitious forces Fω(k−1)
i , Rω(k−1)
i are added to the
external forces F i , Ri . This allows, according to the known solution of the corre-
sponding problem of elasticity theory, to construct a solution to the problem of
viscoplasticity in the recurrent form.
To ensure the convergence of the considered method of successive approxima-
tions, it is necessary that the ω parameter associated with the ϕ(εu ) function of the
relation (10.50) and the non-linear function f (εu ) are small compared to the unit. If
we additionally assume that ϕ (εu )≡ f (εu ), then the specified process of constructing
approximations in images will not differ from the process of constructing approxi-
mations in the method of Ilyushin elastic solutions discussed earlier in the theory of
small elastoplastic deformations.

Remark When constructing solutions to a problem in images, it is convenient for any


(k−1)
approximation to decompose volumetric Fi + Fωi and surface Ri + Rω(k−1)
i forces
into a row according to known functions. In this case, solutions in any approximation
will differ only in constant values.

The method of successive approximations in a form other than the considered


variant is obtained if we take the zero approximation as ω(0) = 0 and ω(0) 1 = 1−
f (0) = 0. In this case, the first approximation ui (1) is defined as a solution to the
usual problem of linear viscoelasticity. For the kth approximation, there is a linear
(k−1)
viscoelastic problem with some other additional “external” loads Fωi , Rω(k−1)
i ,
which are calculated from the results of the previous (k-1)th approximation and are
corrections for the plastic and physical nonlinearity of the material.

Control Questions
1. What is the essence of the method of perturbation of linear elastic properties?
What is the algorithm for solving problems for physically nonlinear elastic
medium?
2. What is the essence of the method of perturbation of linear elastic properties?
What is the algorithm for solving problems for orthotropic media?
3. What is the algorithm for solving the problems of linear theory of viscoelasticity
for inhomogeneous bodies by reducing to solving the problems of the theory of
elasticity of inhomogeneous bodies?
4. What are hardening media?
5. What is the essence of bilinear theory of elasticity?
308 10 Mathematical Models of Special Classes of Solid Mechanics Problems

6. What are the constitutive relations of nonlinear viscoelastic media?


7. What are the Fréchet-Volterre equations of nonlinear viscoelasticity theory?
8. What are the constitutive equations for viscoplastic media?
9. What is the method of successive approximations in viscoplastic problems?

You might also like