100% found this document useful (1 vote)
32 views84 pages

Principles of Continuum Mechanics A Study of Conservation Principles with Applications 1st Edition J. N. Reddy - The full ebook version is just one click away

The document provides information about the ebook 'Principles of Continuum Mechanics' by J. N. Reddy, which serves as an introductory text on mechanics principles for engineering students. It covers fundamental concepts such as stress, strain, and conservation principles, preparing readers for advanced studies in various engineering fields. The ebook is available for download along with other engineering titles on ebookgate.com.

Uploaded by

alanavoelzu7
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
32 views84 pages

Principles of Continuum Mechanics A Study of Conservation Principles with Applications 1st Edition J. N. Reddy - The full ebook version is just one click away

The document provides information about the ebook 'Principles of Continuum Mechanics' by J. N. Reddy, which serves as an introductory text on mechanics principles for engineering students. It covers fundamental concepts such as stress, strain, and conservation principles, preparing readers for advanced studies in various engineering fields. The ebook is available for download along with other engineering titles on ebookgate.com.

Uploaded by

alanavoelzu7
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 84

Get the full ebook with Bonus Features for a Better Reading Experience on ebookgate.

com

Principles of Continuum Mechanics A Study of


Conservation Principles with Applications 1st
Edition J. N. Reddy

https://ebookgate.com/product/principles-of-continuum-
mechanics-a-study-of-conservation-principles-with-
applications-1st-edition-j-n-reddy/

OR CLICK HERE

DOWLOAD NOW

Download more ebook instantly today at https://ebookgate.com


Instant digital products (PDF, ePub, MOBI) available
Download now and explore formats that suit you...

Principles of Engineering Economics with Applications 2nd


Edition Zahid A. Khan

https://ebookgate.com/product/principles-of-engineering-economics-
with-applications-2nd-edition-zahid-a-khan/

ebookgate.com

Principles of Solid Mechanics 1st Edition Rowland Richards


Jr.

https://ebookgate.com/product/principles-of-solid-mechanics-1st-
edition-rowland-richards-jr/

ebookgate.com

Principles of Glacier Mechanics 2nd Edition Roger Leb.


Hooke

https://ebookgate.com/product/principles-of-glacier-mechanics-2nd-
edition-roger-leb-hooke/

ebookgate.com

Principles of Macroeconomics Canadian Edition N. Gregory


Mankiw

https://ebookgate.com/product/principles-of-macroeconomics-canadian-
edition-n-gregory-mankiw/

ebookgate.com
Principles of Economics 6th Edition N. Gregory Mankiw

https://ebookgate.com/product/principles-of-economics-6th-edition-n-
gregory-mankiw/

ebookgate.com

Principles of Economics 10th Edition N. Gregory Mankiw

https://ebookgate.com/product/principles-of-economics-10th-edition-n-
gregory-mankiw/

ebookgate.com

Separation Process Principles with Applications Using


Process Simulators 4th Edition J. D. Seader

https://ebookgate.com/product/separation-process-principles-with-
applications-using-process-simulators-4th-edition-j-d-seader/

ebookgate.com

Continuum and solid mechanics concepts and applications


Victor Quinn

https://ebookgate.com/product/continuum-and-solid-mechanics-concepts-
and-applications-victor-quinn/

ebookgate.com

Principles of Operative Dentistry 1st Edition A. J. E.


Qualtrough

https://ebookgate.com/product/principles-of-operative-dentistry-1st-
edition-a-j-e-qualtrough/

ebookgate.com
This page intentionally left blank
PRINCIPLES OF CONTINUUM MECHANICS
A Study of Conservation Principles with Applications

As most modern technologies are no longer discipline-specific but involve multidis-


ciplinary approaches, undergraduate engineering students should be introduced to
the principles of mechanics so that they have a strong background in the basic prin-
ciples common to all disciplines and are able to work at the interface of science and
engineering disciplines. This textbook is designed for a first course on principles of
mechanics and provides an introduction to the basic concepts of stress and strain and
conservation principles. It prepares engineers and scientists for advanced courses in
traditional as well as emerging fields such as biotechnology, nanotechnology, energy
systems, and computational mechanics. This simple book presents the subjects of
mechanics of materials, fluid mechanics, and heat transfer in a unified form using the
conservation principles of mechanics.

J. N. Reddy is a Distinguished Professor and holder of the Oscar S. Wyatt Endowed


Chair in the Department of Mechanical Engineering at Texas A&M University
(http://www.tamu.edu/acml).
Dr. Reddy is a renowned researcher and educator in the broad fields of mechanics,
applied mathematics, and computational engineering science. Dr. Reddy’s research
areas include theory and finite element analysis of problems in structural mechanics
(composite plates and shells), fluid dynamics, and heat transfer; theoretical model-
ing of stress and deformation of biological cells and soft tissues; nanocomposites;
and development of robust computational technology (including the K-version finite
element models based on the least-squares method in collaboration with Professor
Karan Surana of the University of Kansas). He is the author of more than 400 journal
papers and 16 books on these subjects. His books include Mechanics of Laminated
Plates and Shells: Theory and Analysis, Second Edition, 2004; An Introduction to
Nonlinear Finite Element Analysis, 2004; Introduction to the Finite Element Method,
2006; and An Introduction to Continuum Mechanics, 2008.
Dr. Reddy’s outstanding research credentials have earned him wide international
acclaim in the form of numerous professional awards; citations; fellowship in all
major professional societies including AAM, AIAA, ASC, ASCE, ASME, IACM,
and USACM; membership on two dozen archival journals; and numerous keynote
and plenary lecture invitations at international conferences. Dr. Reddy is the editor-
in-chief of Applied Mechanics Reviews, Mechanics of Advanced Materials and Struc-
tures, International Journal of Computational Methods in Engineering Science and
Mechanics, and International Journal of Structural Stability and Dynamics.
The extent of Dr. Reddy’s original and sustained contributions to education, re-
search, and professional service is substantial. As a result of his extensive publications
of archival journal papers and books on a wide range of topics in applied sciences and
engineering, Dr. Reddy is one of the few researchers in engineering around the world
who is recognized by ISI Highly Cited Researchers, with more than 10,000 citations
with an H-index greater than 46. In February 2009 he was awarded a Honoris Causa
(Honorary Doctorate) by the Technical University of Lisbon.
PRINCIPLES
OF CONTINUUM
MECHANICS
A Study of Conservation
Principles with Applications

J. N. Reddy
Texas A&M University
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore,
São Paulo, Delhi, Dubai, Tokyo

Cambridge University Press


The Edinburgh Building, Cambridge CB2 8RU, UK

Published in the United States of America by Cambridge University Press, New York

www.cambridge.org
Information on this title: www.cambridge.org/9780521513692
© J. N. Reddy 2010

This publication is in copyright. Subject to statutory exception and to the


provision of relevant collective licensing agreements, no reproduction of any part
may take place without the written permission of Cambridge University Press.
First published in print format 2010

ISBN-13 978-0-511-90179-9 eBook (NetLibrary)


ISBN-13 978-0-521-51369-2 Hardback

Cambridge University Press has no responsibility for the persistence or accuracy


of urls for external or third-party internet websites referred to in this publication,
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
When even the brightest mind in our world has been trained up from childhood
in a superstition of any kind, it will never be possible for that mind, in its
maturity, to examine sincerely, dispassionately, and conscientiously any
evidence or any circumstance which shall seem to cast a doubt upon the
validity of that superstition.
Mark Twain

The fact that an opinion has been widely held is no evidence whatever that it
is not utterly absurd; indeed in view of the silliness of the majority of mankind,
a widespread belief is more likely to be foolish than sensible.
Bertrand Russell

Desire for approval and recognition is a healthy motive, but the desire to be
acknowledged as better, stronger, or more intelligent than a fellow being or
fellow scholar easily leads to an excessively egoistic psychological adjustment,
which may become injurious for the individual and for the community.
Albert Einstein
Contents

Preface page xi

1 Introduction 1
1.1 Continuum mechanics 1
1.2 Objective of the study 7
1.3 Summary 8

2 Vectors and Tensors 10


2.1 Motivation 10
2.2 Definition of a vector 10
2.3 Vector algebra 11
2.3.1 Unit vector 11
2.3.2 Zero vector 12
2.3.3 Vector addition 12
2.3.4 Multiplication of a vector by a scalar 14
2.3.5 Scalar product of vectors 14
2.3.6 Vector product 15
2.3.7 Triple products of vectors 18
2.3.8 Plane area as a vector 20
2.3.9 Components of a vector 22
2.4 Index notation and summation convention 24
2.4.1 Summation convention 24
2.4.2 Dummy index 25
2.4.3 Free index 25
2.4.4 Kronecker delta and permutation symbols 26
2.4.5 Transformation law for different bases 29
2.5 Theory of matrices 31
2.5.1 Definition 31
2.5.2 Matrix addition and multiplication of a matrix by a scalar 32
2.5.3 Matrix transpose and symmetric and skew symmetric matrices 33
2.5.4 Matrix multiplication 34
2.5.5 Inverse and determinant of a matrix 36

vii
viii Contents

2.6 Vector calculus 39


2.6.1 The del operator 39
2.6.2 Divergence and curl of a vector 41
2.6.3 Cylindrical and spherical coordinate systems 43
2.6.4 Gradient, divergence, and curl theorems 45
2.7 Tensors 46
2.7.1 Dyads 46
2.7.2 Nonion form of a dyad 48
2.7.3 Transformation of components of a dyad 49
2.7.4 Tensor calculus 49
2.8 Summary 51
Problems 51

3 Kinematics of a Continuum 55
3.1 Deformation and configuration 55
3.2 Engineering strains 56
3.2.1 Normal strain 56
3.2.2 Shear strain 57
3.3 General kinematics of a solid continuum 61
3.3.1 Configurations of a continuous medium 61
3.3.2 Material and spatial descriptions 62
3.3.3 Displacement field 65
3.4 Analysis of deformation 66
3.4.1∗ Deformation gradient tensor 66
3.4.2∗ Various types of deformations 69
3.4.2.1 Pure dilatation 70
3.4.2.2 Simple extension 70
3.4.2.3 Simple shear 71
3.4.2.4 Nonhomogeneous deformation 71
3.4.3 Green strain tensor 72
3.4.4 Infinitesimal strain tensor 77
3.4.5 Principal values and principal planes of strains 79
3.5 Rate of deformation and vorticity tensors 81
3.5.1 Velocity gradient tensor 81
3.5.2 Rate of deformation tensor 81
3.5.3 Vorticity tensor and vorticity vector 82
3.6 Compatibility equations 84
3.7 Summary 86
Problems 87
4 Stress Vector and Stress Tensor 93
4.1 Introduction 93
4.2 Stress vector, stress tensor, and Cauchy’s formula 94
4.3 Transformations of stress components and principal stresses 102
ix Contents

4.3.1 Transformation of stress components 102


4.3.2 Principal stresses and principal planes 104
4.4 Summary 107
Problems 107

5 Conservation of Mass, Momentum, and Energy 111


5.1 Introduction 111
5.2 Conservation of mass 112
5.2.1 Preliminary discussion 112
5.2.2 Conservation of mass in spatial description 112
5.2.3 Conservation of mass in material description 117
5.2.4 Reynolds transport theorem 119
5.3 Conservation of momenta 119
5.3.1 Principle of conservation of linear momentum 119
5.3.2 Principle of conservation of angular momentum 134
5.4 Thermodynamic principles 136
5.4.1 Introduction 136
5.4.2 Energy equation for one-dimensional flows 136
5.4.3 Energy equation for a three-dimensional continuum 140
5.5 Summary 142
Problems 143

6 Constitutive Equations 149


6.1 Introduction 149
6.2 Elastic solids 150
6.2.1 Introduction 150
6.2.2 Generalized Hooke’s law for orthotropic materials 151
6.2.3 Generalized Hooke’s law for isotropic materials 153
6.3 Constitutive equations for fluids 156
6.3.1 Introduction 156
6.3.2 Ideal fluids 157
6.3.3 Viscous incompressible fluids 157
6.4 Heat transfer 158
6.4.1 General introduction 158
6.4.2 Fourier’s heat conduction law 158
6.4.3 Newton’s law of cooling 159
6.4.4 Stefan–Boltzmann law 159
6.5 Summary 160
Problems 160
7 Applications in Heat Transfer, Fluid Mechanics, and Solid Mechanics 162
7.1 Introduction 162
7.2 Heat transfer 162
7.2.1 Governing equations 162
x Contents

7.2.2 Analytical solutions of one-dimensional heat transfer 165


7.2.2.1 Steady-state heat transfer in a cooling fin 165
7.2.2.2 Steady-state heat transfer in a surface-insulated rod 167
7.2.3 Axisymmetric heat conduction in a circular cylinder 169
7.2.4 Two-dimensional heat transfer 170
7.3 Fluid mechanics 172
7.3.1 Preliminary comments 172
7.3.2 Summary of equations 173
7.3.3 Inviscid fluid statics 174
7.3.4 Parallel flow (Navier–Stokes equations) 175
7.3.4.1 Steady flow of viscous incompressible fluid between
parallel plates 176
7.3.4.2 Steady flow of viscous incompressible fluid through a
pipe 177
7.3.5 Diffusion processes 179
7.4 Solid mechanics 182
7.4.1 Governing equations 182
7.4.2 Analysis of bars 184
7.4.3 Analysis of beams 188
7.4.3.1 Principle of superposition 195
7.4.4 Analysis of plane elasticity problems 196
7.4.4.1 Plane strain and plane stress problems 196
7.4.4.2 Plane strain problems 196
7.4.4.3 Plane stress problems 198
7.4.4.4 Solution methods 199
7.4.4.5 Airy stress function 202
7.5 Summary 204
Problems 205

Answers to Selected Problems 215


References and Additional Readings 225
Subject Index 227
Preface

You cannot teach a man anything, you can only help him find it within himself.
Galileo Galilei

This book is a simplified version of the author’s book, An Introduction to Con-


tinuum Mechanics with Applications, published by Cambridge University Press
(New York, 2008), intended for use as an undergraduate textbook. As most mod-
ern technologies are no longer discipline-specific but involve multidisciplinary
approaches, undergraduate engineering students should be educated to think and
work in such environments. Therefore, it is necessary to introduce the subject of
principles of mechanics (i.e., laws of physics applied to science and engineer-
ing systems) to undergraduate students so that they have a strong background
in the basic principles common to all disciplines and are able to work at the
interface of science and engineering disciplines. A first course on principles of
mechanics provides an introduction to the basic concepts of stress and strain
and conservation principles and prepares engineers and scientists for advanced
courses in traditional as well as emerging fields such as biotechnology, nanotech-
nology, energy systems, and computational mechanics. Undergraduate students
with such a background may seek advanced degrees in traditional (e.g., aerospace,
civil, electrical or mechanical engineering; physics; applied mathematics) as well
as interdisciplinary (e.g., bioengineering, engineering physics, nanoscience and
engineering, biomolecular engineering) degree programs.
There are not many books on principles of mechanics that are written that
keep the undergraduate engineering or science student in mind. A vast majority
of books on the subject are written for graduate students of engineering and tend
to be more mathematical and too advanced to be of use for third-year or senior
undergraduate students. This book presents the subjects of mechanics of mate-
rials, fluid mechanics, and heat transfer in unified form using the conservation
principles of mechanics. It is hoped that the book, which is simple, will facilitate
in presenting the main concepts of the previous three courses under a unified
framework.
With a brief discussion of the concept of a continuum in Chapter 1, a review of
vectors and tensors is presented in Chapter 2. Because the analytical language of
applied sciences and engineering is mathematics, it is necessary for all students

xi
xii Preface

of this course to familiarize themselves with the notation and operations of


vectors, matrices, and tensors that are used in the mathematical description of
physical phenomena. Readers who are familiar with the topics of this chapter may
refresh or skip and go to the next chapter. The subject of kinematics, which deals
with geometric changes without regard to the forces causing the deformation,
is discussed in Chapter 3. Measures of engineering normal and shear strains
and definitions of mathematical strains are introduced here. Both simple one-
dimensional systems as well as two-dimensional continua are used to illustrate
the strain and strain-rate measures introduced. In Chapter 4, the concept of stress
vector and stress tensor are introduced. It is here that the readers are presented
with entities that require two directions – namely, the plane on which they are
measured and the direction in which they act – to specify them. Transformation
equations among components of stress tensor referred to two different orthogonal
coordinate systems are derived, and principal values and principal planes (i.e.,
eigenvalue problems associated with the stress tensor) are also discussed.
Chapter 5 is dedicated to the derivation of the governing equations of mechanics
using the conservation principles of continuum mechanics (or laws of physics).
The principles of conservation of mass, linear momentum, angular momentum,
and energy are presented using one-dimensional systems as well as general three-
dimensional systems. The derivations are presented in invariant (i.e., independent
of a coordinate system) as well as in component form. The equations resulting
from these principles are those governing stress and deformation of solid bodies,
stress and rate of deformation of fluid elements, and transfer of heat through solid
media. Thus, this chapter forms the heart of the course. Constitutive relations that
connect the kinematic variables (e.g., density, temperature, deformation) to the
kinetic variables (e.g., internal energy, heat flux, stresses) are discussed in Chapter
6 for elastic materials, viscous fluids, and heat transfer in solids.
Chapter 7 is devoted to the application of the field equations derived in Chapter
5 and constitutive models presented in Chapter 6 to problems of heat conduction
in solids, fluid mechanics (inviscid flows as well as viscous incompressible flows),
diffusion, and solid mechanics (e.g., bars, beams, and plane elasticity). Simple
boundary-value problems are formulated and their solutions are discussed. The
material presented in this chapter illustrates how physical problems are analyt-
ically formulated with the aid of the equations resulting from the conservation
principles.
As stated previously, the present book is an undergraduate version of the au-
thor’s book An Introduction to Continuum Mechanics (Cambridge University
Press, New York, 2008). The presentation herein is limited in scope when com-
pared to the author’s graduate-level textbook. The major benefit of a course based
on this book is to present the governing equations of diverse physical phenomena
from a unified point of view, namely, from the conservation principles (or laws
of physics), so that students of applied science and engineering see the physi-
cal principles as well as the mathematical structure common to diverse fields.
xiii Preface

Readers interested in advanced topics may consult the author’s continuum me-
chanics book or other titles listed in references therein.
The author is pleased to acknowledge the fact that the manuscript was tested
with the undergraduate students in the College of Engineering at Texas A&M
University as well as in the Engineering Science Programme at the National
University of Singapore. The students, in general, have liked the contents and
the simplicity with which the concepts are introduced and explained. They also
expressed the feeling that the subject is more challenging than most at the under-
graduate level but a useful prerequisite to graduate courses in engineering.
The author wishes to thank Drs. Vinu and Ginu Unnikrishnan and Ms. Feifei
Cheng for their help with the proofreading of the manuscript of this book dur-
ing the course of its preparation and production. The book contains so many
mathematical expressions that it is hardly possible not to have typographical and
other kinds of errors. The author wishes to thank in advance those who are will-
ing to draw the author’s attention to typos and errors, using the e-mail address
jn [email protected].
1 Introduction

One thing I have learned in a long life: that all our science, measured against
reality, is primitive and childlike—and yet it is the most precious thing we have.
Albert Einstein

1.1 Continuum mechanics

Matter is composed of discrete molecules, which in turn are made up of atoms. An


atom consists of electrons, positively charged protons, and neutrons. Electrons
form chemical bonds. An example of mechanical (i.e., has no living cells) matter
is a carbon nanotube (CNT), which consists of carbon molecules in a certain
geometric pattern in equilibrium with each other, as shown in Figure 1.1.1.
Another example of matter is a biological cell, which is a fundamental unit
of any living organism. There are two types of cells: prokaryotic and eukaryotic
cells. Eukaryotic cells are generally found in multicellular organs and have a
true nucleus, distinct from a prokaryotic cell. Structurally, cells are composed of
a large number of macromolecules, or large molecules. These macromolecules
consist of large numbers of atoms and form specific structures, like chromosomes
and plasma membranes in a cell. Macromolecules occur as four major types:
carbohydrates, proteins, lipids, and nucleic acids. To highlight the hierarchical
nature of the structures formed by the macromolecule in a cell, let us analyze a
chromosome.
Chromosomes, which are carriers of hereditary traits in an individual, are
found inside the nucleus of all eukaryotes. Each chromosome consists of a sin-
gle nucleic acid macromolecule called deoxyribonucleic acid (DNA), each 2.2–
2.4 nanometers wide. These nucleic acids are in turn formed from the specific
arrangement of monomers called mononucleotides, each 0.3–0.33 nanometers
wide. The fundamental units of nucleotides are formed again by a combination
of a specific arrangement of a phosphate radical, nitrogenous base, and a carbo-
hydrate sugar. The hierarchical nature of the chromosome is shown in Fig-
ure 1.1.2(a). Similar to the chromosomes, all the structures in a cell are formed
from a combination of the macromolecules.

1
2 Principles of Continuum Mechanics

(a) (b)
STRIP OF A GRAPHENE SHEET ROLLED INTO A TUBE

(n,0) / ZIG ZAG

(m,m) / ARM CHAIR

CHIRAL
(m,n)

Carbon nanotubes (CNTs) with different chiralities.


Figure 1.1.1

At the macroscopic scale, eukaryotic cells can be divided into three distinct
regions: nucleus, plasma membrane, and a cytoplasm having a host of other
structures, as shown in Figure 1.1.2(b). The nucleus consists of the chromosomes
and other protein structures and is the control center of the cell determining
how the cell functions. The plasma membrane encloses the cell and separates
the material outside the cell from inside. It is responsible for maintaining the
integrity of the cell and also acts as channels for the transport of molecules to and
from the cell. The cell membrane is made up of a double layer of phospholipid
molecules (macromolecules) having embedded transmembrane proteins. The re-
gion between the cell membrane and the nucleus is the cytoplasm, which consists
of a gel-like fluid called cytosol, the cytoskeleton, and other macromolecules.
The cytoskeleton forms the biomechanical framework of the cell and consists
of three primary protein macromolecule structures of actin filaments, intermedi-
ate filaments, and microtubules. Growth, cell expansion, and replication are all
carried out in the cytoplasm.
The interactions between the different components of the cell are responsible
for maintaining the structural integrity of the cell. The analysis of these interac-
tions to obtain the response of the cell when subjected to an external stimulus
(mechanical, electrical, or chemical) is studied systematically under cell mechan-
ics. The structural framework of primary macromolecular structures in a cell is
shown in Figure 1.1.2(c).
The study of matter at molecular or atomistic levels is very useful for under-
standing a variety of phenomena, but studies at these scales are not useful to
solve common engineering problems. The understanding gained at the molecular
3 Introduction

Nitrogen Base

Phosphate

DNA (2.2–2.4 nm) Nucleotides (0.3–0.33 nm)


(a)

Cell Membrane Nucleus


Cytoplasm Organelles
(interior contents
of the cell)

Actin filaments
Intermediate filaments Network of actin filaments
Microtubule
Cytosol – Heterogeneous fluid component
Cytoskeleton – Filament network permeating the cell’s interior
(b)
(outside of cell)

membrane protein

glycoprotein
cholesterol

phospholipid

plasma
membrane

(cytoplasm inside of cell)

Cell Membrane
Plasma
Membrane
Animal Cell
Endoplasmic Microtubule
Reticulum
Mitochondrion

Ribosomes

Microfilaments
Microfilaments and Intermediate
Filaments

Microtubule

F-Action G-Action
(Polymer) (Monomer)

(c)
(a) Hierarchical nature of a chromosome. (b) Structure of a generalized cell. (c) Macromolecular structure in a
Figure 1.1.2
cell.
4 Principles of Continuum Mechanics

level must be taken to the macrosopic


scale (i.e., a scale that a human eye can
see) to be able to study its behavior.
Central to this study is the assump-
tion that the discrete nature of mat-
ter can be overlooked, provided the
length scales of interest are large com-
pared to the length scales of discrete
molecular structure. Thus, matter at
sufficiently large length scales can be
Progressive damage of artery due to deposition of par-
Figure 1.1.3
ticles in the arterial wall. treated as a continuum in which all
physical quantities of interest, includ-
ing density, are continuously differentiable.
The subject of mechanics deals with the study of motion and forces in solids,
liquids, and gases and the deformation or flow of these materials. In such a
study, we make the simplifying assumption, for analytic purposes, that the matter
is distributed continuously, without gaps or empty spaces (i.e., we disregard
the molecular structure of matter). Such a hypothetical continuous matter is
termed a continuum. In essence, in a continuum all quantities such as the density,
displacements, velocities, stresses, and so on vary continuously so that their
spatial derivatives exist and are continuous. The continuum assumption allows
us to shrink an arbitrary volume of material to a point, in much the same way
as we take the limit in defining a derivative, so that we can define quantities of
interest at a point. For example, the density (mass per unit volume) of a material
at a point is defined as the ratio of the mass m of the material to a small volume
V surrounding the point in the limit that V becomes a value  3 , where  is
small compared with the mean distance between molecules,
m
ρ = lim . (1.1.1)
V → 3 V
In fact, we take the limit  → 0. A mathematical study of mechanics of such an
idealized continuum is called continuum mechanics.
Engineers and scientists undertake the study of continuous systems to un-
derstand their behavior under “working conditions,” so that the systems can be
designed to function properly and produced economically. For example, if we
were to repair or replace a damaged artery in a human body, we must understand
the function of the original artery and the conditions that lead to its damage. An
artery carries blood from the heart to different parts of the body. Conditions like
high blood pressure and increases in cholesterol content in the blood may lead to
deposition of particles in the arterial wall, as shown in Figure 1.1.3. With time,
accumulation of these particles in the arterial wall hardens and constricts the
passage, leading to cardiovascular diseases. A possible remedy for such diseases
is to repair or replace the damaged portion of the artery. This in turn requires an
understanding of the deformation and stresses caused in the arterial wall by the
5 Introduction

A diving board fixed at the left end and free at the right end.
Figure 1.1.4

flow of blood. The understanding is then used to design the vascular prosthesis
(i.e., an artificial artery).
The primary objectives of this book are (1) to study the conservation principles
in mechanics of continua and formulate the equations that describe the motion
and mechanical behavior of materials, and (2) to present the applications of
these equations to simple problems associated with flows of fluids, conduction
of heat, and deformation of solid bodies. Although the first of these objectives is
an important topic, the reason for the formulation of the equations is to gain a
quantitative understanding of the behavior of an engineering system. This quanti-
tative understanding is useful in the design and manufacture of better products.
Typical examples of engineering problems sufficiently simple to cover in this
course are described in the following. At this stage of discussion, it is sufficient
to rely on the reader’s intuitive understanding of concepts.

PROBLEM 1 (MECHANICAL STRUCTURE)


We wish to design a diving board that must enable the swimmer to gain enough
momentum for the swimming exercise. The diving board is fixed at one end
and free at the other end (see Figure 1.1.4). The board is initially straight and
horizontal, and of length L and uniform cross section A = bh.
The design process consists of selecting the material with Young’s modulus
E and cross-sectional dimensions b and h such that the board carries the weight
W of the swimmer. The design criteria are that the stresses developed do not
exceed the allowable stress and the deflection of the free end does not exceed
a pre-specified value δ. A preliminary design of such systems is often based
on mechanics of materials equations. The final design involves the use of more
sophisticated equations, such as the three-dimensional elasticity equations. The
equations of elementary beam theory may be used to find a relation between the
deflection δ of the free end in terms of the length L, cross-sectional dimensions
b and h, Young’s modulus E, and weight W :
4W L 3
δ= . (1.1.2)
Ebh 3
6 Principles of Continuum Mechanics

r
vx(r )
Internal diameter, d
x P1 P2
L
Measurement of viscosity of a fluid using a capillary tube.
Figure 1.1.5

Given δ (allowable deflection) and load W (maximum possible weight of a


swimmer), one can select the material (Young’s modulus, E) and dimensions
L, b, and h (which must be restricted to the standard sizes fabricated by a
manufacturer). In addition to the deflection criterion, one must also check if
the board develops stresses that exceed the allowable stresses of the material
selected. Analysis of pertinent equations provide the designer with alternatives
to select the material and dimensions of the board so as to have a cost-effective
but functionally reliable structure.

PROBLEM 2 (FLOW OF FLUIDS)


We wish to measure the viscosity µ of a lubricating oil used in rotating machinery
to prevent the damage of the parts in contact. Viscosity, like Young’s modulus
of solid materials, is a material property that is useful in the calculation of shear
stresses developed between a fluid and a solid body.
A capillary tube is used to determine the viscosity of a fluid via the formula
π d 4 P1 − P2
µ= , (1.1.3)
128L Q
where d is the internal diameter and L is the length of the capillary tube, P1 and
P2 are the pressures at the two ends of the tube (oil flows from one end to the
other, as shown in Figure 1.1.5), and Q is the volume rate of flow at which the
oil is discharged from the tube. As we shall see later in this course, Eq. (1.1.3) is
derived using the principles of continuum mechanics.

PROBLEM 3 (TRANSFER OF HEAT IN SOLIDS)


We wish to determine the heat loss through the wall of a furnace. The wall typically
consists of layers of brick, cement mortar, and cinder block (see Figure 1.1.6).
Each of these materials provides varying degrees of thermal resistance. The
Fourier heat conduction law,
dT
, q = −k (1.1.4)
dx
provides a relation between the heat flux q (heat flow per unit area) and gra-
dient of temperature T . Here k denotes thermal conductivity (1/k is the ther-
mal resistance) of the material. The negative sign in Eq. (1.1.4) indicates that
heat flows from a high-temperature region to a low-temperature region. Using
the continuum mechanics equations, one can determine the heat loss when the
7 Introduction

Cross section
of the wall

x
Furnace
Heat transfer through a composite wall of a furnace.
Figure 1.1.6

temperatures inside and outside of the building are known. A building designer
can select the materials as well as thicknesses of various components of the wall
to reduce the heat loss while ensuring necessary structural strength – a structural
analysis aspect.
The previous three examples provide some indication of the need for studying
the response of materials under the influence of external loads. The response of
a material is consistent with the laws of physics and the constitutive behavior of
the material. This book has the objective of describing the physical principles
and deriving the equations governing the stress and deformation of continuous
materials, and then solving some simple problems from various branches of
engineering to illustrate the applications of the principles discussed and equations
derived.

1.2 Objective of the study

The primary objective of this book, as already stated, is twofold: (1) use the
physical principles to derive the equations that govern the motion and thermome-
chanical response of materials and systems, and (2) apply these equations for the
solution of specific problems of engineering and applied science (e.g., linearized
elasticity, heat transfer, and fluid mechanics). The governing equations for the
study of deformation and stress of a continuous material are nothing but an ana-
lytical representation of the global laws of conservation of mass, momenta, and
energy, and the constitutive response of the continuum. They are applicable to
all materials that are treated as a continuum. Tailoring these equations to partic-
ular problems and solving them constitute the bulk of engineering analysis and
design.
The study of motion and deformation of a continuum (or a “body” consisting
of continuously distributed material) can be broadly classified into four basic
categories:

(1) Kinematics
(2) Kinetics (conservation of linear and angular momentum)
(3) Thermodynamics (first and second laws of thermodynamics)
(4) Constitutive equations
8 Principles of Continuum Mechanics

Table 1.2.1. Four major topics of the present study, principles of mechanics used, resulting governing equations, and
variables involved.
Topic of study Physical principle Resulting equations Variables involved
1. Kinematics Based on geometric Strain-displacement relations Displacements and strains
changes
Strain rate-velocity relations Velocities and strain rates
2. Kinetics Conservation of linear Equations of motion Stresses, velocities, and body
momentum forces
Conservation of angular Symmetry of stress tensor Stresses
momentum
3. Thermodynamics First law Energy equation Temperature, heat flux,
stresses, heat generation,
and velocities
Second law Clausius–Duhem inequality Temperature, heat flux, and
entropy
4. Constitutive equations Constitutive axioms Hooke’s law Stresses, strains, heat flux,
(not all relations are and temperature
listed)
Newtonian fluids Stresses, pressure, velocities
Fourier’s law Heat flux and temperature
Equations of state Density, pressure,
temperature
5. Boundary conditions All of the previous Relations between kinematic All of the previous variables
principles and axioms and kinetic variables

Kinematics is the study of the geometric changes or deformation in a contin-


uum, without the consideration of forces causing the deformation. Kinetics is the
study of the static or dynamic equilibrium of forces and moments acting on a
continuum, using the principles of conservation of linear and angular momen-
tum. This study leads to equations of motion as well as the symmetry of stress
tensors in the absence of body couples. Thermodynamic principles are concerned
with the conservation of energy and relations among heat, mechanical work, and
thermodynamic properties of the continuum. Constitutive equations describe the
thermomechanical behavior of the material of the continuum and relate the de-
pendent variables introduced in the kinetic description to those introduced in
the kinematic and thermodynamic descriptions. Table 1.2.1 provides a brief sum-
mary of the relationship between physical principles and governing equations, and
physical entities involved in the equations. To the equations derived from physical
principles, one must add boundary conditions of the system (and initial conditions
if the phenomenon is time-dependent) to complete the analytical description.

1.3 Summary

In this chapter, the concept of a continuous medium is discussed with the major
objectives of the present study, namely, to use the principles of mechanics to
9 Introduction

derive the equations governing a continuous medium and to present applications


of the equations in the solution of specific problems arising in engineering. The
study of principles of mechanics is broadly divided into four topics, as outlined in
Table 1.2.1. These four topics form the subject of Chapters 3 through 6, respec-
tively. Mathematical formulation of the governing equations of a continuous
medium (that is, the development of a mathematical model of the physical phe-
nomenon) necessarily requires the use of vectors, matrices, and tensors – mathe-
matical tools that facilitate analytical formulation of the natural laws. Therefore,
it is useful to first gain certain operational knowledge of vectors, matrices, and
tensors. Chapter 2 is dedicated to this purpose.
Many of the concepts presented herein are the same as those most likely in-
troduced in undergraduate courses on mechanics of materials, heat transfer, fluid
mechanics, and material science. The present course brings together these courses
under a common mathematical framework and, thus, may require mathematical
tools as well as concepts not seen previously. The readers must motivate and
challenge themselves to learn the new mathematical concepts introduced here, as
the language of engineers is mathematics. This subject also serves as a prelude
to many graduate courses in engineering and applied sciences.
Although this book is self-contained for an introduction to principles of contin-
uum mechanics, there are several books that may provide an advanced treatment
of the subject. The graduate-level textbook by the author, An Introduction to Con-
tinuum Mechanics with Applications (Cambridge University Press, New York,
2008), provides additional material. Interested readers may consult other titles
listed in “References and Additional Readings,” at the end of this book.

When a distinguished but elderly scientist states that something is possible, he


is almost certainly right. When he states that something is impossible, he is very
probably wrong.
Arthur C. Clarke
2 Vectors and Tensors

No great discovery was ever made without a bold guess.


Isaac Newton

2.1 Motivation

In the mathematical description of equations governing a continuous medium,


we derive relations between various quantities that describe the response of the
continuum by means of the laws of nature, such as Newton’s laws. As a means of
expressing a natural law, a coordinate system in a chosen frame of reference is
often introduced. The mathematical form of the law thus depends upon the chosen
coordinate system and may appear different in another coordinate system. How-
ever, the laws of nature should be independent of the choice of coordinate system,
and we may seek to represent the law in a manner independent of a particular
coordinate system.1 A way of doing this is provided by objects called vectors and
tensors. When vector and tensor notation is used, a particular coordinate system
need not be introduced. Consequently, the use of vector and tensor notation in
formulating natural laws leaves them invariant, and we may express them in any
chosen coordinate system. A study of physical phenomena by means of vectors
and tensors can lead to a deeper understanding of the problem, in addition to
bringing simplicity and versatility to the analysis. This chapter is dedicated to the
algebra and calculus of physical vectors and tensors, as needed in the subsequent
study.

2.2 Definition of a vector

The quantities encountered in the analytical description of physical phenomena


can be classified into the following groups according to the information needed
to specify them completely:
1 We always return to a particular coordinate system of our choice to solve the equations resulting
from the physical law.

10
11 Vectors and Tensors

.
Magnitude (length)
of the vector, A
A = A êA

. êA

Origin of the vector


Geometric representation of a physical vector.
Figure 2.2.1

Scalars Nonscalars
Mass Force
Temperature Moment
Time Stress
Volume Acceleration
Length Displacement

The scalars are given by a single number. Nonscalars not only have a magnitude
specified but also have additional information, such as direction. Nonscalars that
obey certain rules (such as the parallelogram law of addition) are called vectors.
Not all nonscalar quantities are vectors, unless they obey certain rules as discussed
in the sequel.
A physical vector is often shown as a directed line segment with an arrow head
at the end of the line, as shown in Figure 2.2.1. The length of the line represents
the magnitude of the vector and the arrow indicates the direction. In written or
typed material, it is customary to place an arrow over the letter denoting the
 In printed material, the letter used for the vector is commonly
vector, such as A.
denoted by a boldface letter, A, such as used in this study. The magnitude of the
vector A is denoted by |A|, A, or A. The magnitude of a vector is a scalar.

2.3 Vector algebra

In this section, we discuss various operations with vectors and interpret them
physically. First, we introduce the notion of unit and zero vectors.

2.3.1 Unit vector


A vector of unit length is called a unit vector. The unit vector along A can be
defined as follows:
A
ê A = . (2.3.1)
A
We can now write

A = A ê A . (2.3.2)
12 Principles of Continuum Mechanics

A
A+B
B

A B+A

+ = = A
A

B B
B+A
A

B
(a) (b)
(a) Addition of vectors. (b) Parallelogram law of addition.
Figure 2.3.1

Thus, any vector can be represented as a product of its magnitude and a unit
vector. A unit vector is used to designate direction. It does not have any phys-
ical dimensions. We denote a unit vector by a “hat” (caret) above the boldface
letter, ê.

2.3.2 Zero vector


A vector of zero magnitude is called a zero vector or a null vector. All null vectors
are considered equal to each other without consideration as to their direction. Note
that a lightface zero, 0, is a scalar and a boldface zero, 0, is the zero vector.

2.3.3 Vector addition


Let A, B, and C be any vectors. Then there exists a vector A+B, called the sum
of A and B, such that:
(1) A + B = B + A (commutative property).
(2) (A + B) + C = A + (B + C) (associative property).
(3) There exists a unique vector, 0, independent of A such that
A + 0 = A (existence of zero vector). (2.3.3)
(4) To every vector A, there exists a unique vector −A (that depends on A)
such that
A + (−A) = 0 (existence of negative vector).
The addition of two vectors is shown in Figure 2.3.1(a). Note that the commutative
property is essential for a nonscalar to qualify as a vector. The combination of
the two diagrams in Figure 2.3.1(a) gives the parallelogram shown in Figure
2.3.1(b), and it characterizes the commutativity. Thus, we say that vectors add
13 Vectors and Tensors

R = θ n̂
Defined direction (thumb)
y x
(with double arrow)
Sense of rotation
x^y

(a)

z Final position
y A A
A

A
x
Rotation 1 Rotation 2
+90 deg about the z-axis −90 deg about the y-axis
(b)

Final position

A
A

A
A

Rotation 1 Rotation 2
−90 deg about the y-axis +90 deg about the z-axis
(c)
(a) Preferred sense of rotation. (b) Rotation θz followed by rotation θ y . (c) Rotation θ y followed by rotation θz .
Figure 2.3.2

according to the parallelogram law of addition. The negative vector −A has the
same magnitude as A but the opposite sense. Subtraction of vectors is carried out
along the same lines. To form the difference A − B, we write

A − B = A + (−B) (2.3.4)

and subtraction reduces to the operation of addition.


As an example of a nonscalar that has magnitude and direction but does not
obey commutativity, consider finite rotation. Finite rotation has a magnitude θ
and a preferred direction as that in which a right-handed screw would advance
when turned in the direction of rotation, as indicated in Figure 2.3.2(a). Now
consider two different rotations of a rectangular block, in a certain order. The
first rotation is about the z-axis by θz = +90◦ , followed by rotation θ y = −90◦
about the y-axis. This sequence of rotations results in the final position indicated
in Figure 2.3.2(b). We may represent this pair of rotations as R1 + R2 , as shown
in Figure 2.3.2(c). Reversing the order of rotations, that is, θ y first and θz next, we
obtain R2 + R1 , which is not the same as the position achieved by R1 + R2 . Thus,
a finite rotation is not a vector, even though it has a direction and magnitude.
14 Principles of Continuum Mechanics

A αA (α > 1)
αA (α < 1)

A typical vector A and its scalar multiple.


Figure 2.3.3

2.3.4 Multiplication of a vector by a scalar


Let A and B be vectors and α and β be real numbers (scalars). To every vector A
and every real number α, there correspond a unique vector αA such that:
(1) α(βA) = (αβ)A (associative property).
(2) (α + β)A = αA + βA (distributive scalar addition).
(2.3.5)
(3) α(A + B) = αA + αB (distributive vector addition).
(4) 1 · A = A · 1 = A, 0 · A = 0.
Figure 2.3.3 contains a vector A and its scalar multiple αA for α > 1 and α < 1.
Two vectors A and B are equal if their magnitudes are equal, |A| = |B|, and
if their directions are equal. Consequently, a vector is not changed if it is moved
parallel to itself. This means that the position of a vector in space – that is, the
point from which the line segment is drawn (or the end without the arrowhead) –
may be chosen arbitrarily. However, in certain applications the actual point of
the location of a vector may be important, for instance, for a moment or a force
acting on a body. A vector associated with a given point is known as a localized
or bound vector.
Two vectors A and B are said to be linearly dependent if they are scalar multiples
of each other, that is, c1 A + c2 B = 0 for some nonzero scalars c1 and c2 . If two
vectors are linearly dependent, then they are collinear. If three vectors A, B, and C
are linearly dependent, then they are coplanar. A set of n vectors {A1 , A2 , . . . , An }
is said to be linearly dependent if a set of n numbers c1 , c2 , . . . , cn can be found
such that c1 A1 + c2 A2 + · · · + cn An = 0, where c1 , c2 , . . . , cn cannot all be zero.
If this expression cannot be satisfied (that is, all ci are zero), the set of vectors
{A1 , A2 , . . . , An } is said to be linearly independent.

2.3.5 Scalar product of vectors


When a force F acts on a point mass and moves through a displacement vector
d, as shown in Figure 2.3.4(a), the work done by the force vector is defined
by the projection of the force in the direction of the displacement, as shown
in Figure 2.3.4(b), multiplied by the magnitude of the displacement. Such an
operation can be defined for any two vectors. Because the result of the product is
a scalar, it is called the scalar product. We denote this product as F · d ≡ (F, d),
and it is defined as follows:

F · d ≡ (F, d) = Fd cos θ, 0 ≤ θ ≤ π. (2.3.6)


The scalar product is also known as the dot product or inner product.
15 Vectors and Tensors

F
θ

(a)

Projection of vector F
F onto vector d
θ
d

(b)
(a) Representation of work done. (b) Projection of a vector.
Figure 2.3.4
A few simple results follow from the definition in Eq. (2.3.6), and they are
listed next.

(1) Because A · B = B · A, the scalar product is commutative.


(2) If the vectors A and B are perpendicular to each other, then A · B =
AB cos(π/2) = 0. Conversely, if A · B = 0, then either A or B is zero or
A is perpendicular, or orthogonal, to B.
(3) If two vectors A and B are parallel and in the same direction, then A · B =
AB cos 0 = AB, because cos 0 = 1. Thus, the scalar product of a vector
with itself is equal to the square of its magnitude:

A · A = A A = A2 . (2.3.7)

(4) The orthogonal projection of a vector A in any direction ê is given by


(A · ê)ê.
(5) The scalar product follows the distributive law:

A · (B + C) = (A · B) + (A · C). (2.3.8)

2.3.6 Vector product


Consider the concept of moment due to a force. Let us describe the moment about
a point O of a force F acting at a point P, such as that shown in Figure 2.3.5(a).

+
F r×F
F

O r θ
θ
P O
l r
(a) (b)
(a) Representation of a moment. (b) Direction of rotation.
Figure 2.3.5
16 Principles of Continuum Mechanics

F
M êM
θ

r
Axis of rotation.
Figure 2.3.6

By definition, the magnitude of the moment is given by

M=F , F = |F|, (2.3.9)

where is the perpendicular distance from the point O to the force F (called the
lever arm). If r denotes the vector OP and θ the angle between r and F, as shown
in Figure 2.3.5(a), such that 0 ≤ θ ≤ π, we have = r sin θ, and thus

M = Fr sin θ. (2.3.10)

A direction can now be assigned to the moment. Drawing the vectors F and
r from the common origin O, we note that the rotation due to F tends to bring
r into F, as can be seen from Figure 2.3.5(b). We now set up an axis of rotation
perpendicular to the plane formed by F and r. Along this axis of rotation, we set up
a preferred direction as that in which a right-handed screw would advance when
turned in the direction of rotation due to the moment, as shown in Figure 2.3.6.
Along this axis of rotation, we draw a unit vector ê M and agree that it represents
the direction of the moment M. Thus, we have

M = M ê M = Fr sin θ ê M (2.3.11)
= r × F. (2.3.12)

According to this expression, M can be looked upon as resulting from a special


operation between the two vectors F and r. Thus, it is the basis for defining a
product between any two vectors. Because the result of such a product is a vector,
it is called the vector product.
The product of two vectors A and B is a vector C whose magnitude is equal
to the product of the magnitudes of vectors A and B multiplied by the sine of the
angle measured from A to B such that 0 ≤ θ ≤ π . The direction of vector C is
specified by the condition that C be perpendicular to the plane of the vectors A
and B, and point in the direction in which a right-handed screw advances when
turned so as to bring A into B, as shown in Figure 2.3.7.
The vector product is usually denoted by

C = A × B = AB sin(A, B) ê = AB sin θ ê, (2.3.13)

where sin(A,B) denotes the sine of the angle between vectors A and B. This
product is called the cross product or vector product. When A = a ê A and B =
b ê B are the vectors representing the sides of a parallelogram, with a and b
denoting the lengths of the sides, then the magnitude of the vector product A × B
17 Vectors and Tensors

B sin θ
A×B
B
B
C=A×B ê θ A sin θ θ

A A
A × B = −B × A
−B ×A
Representation of the vector product.
Figure 2.3.7
represents the area of the parallelogram, ab sin θ. The unit vector ê = ê A × ê B
denotes the normal to the plane area. Thus, an area can be represented as a vector
(see Section 2.3.8 for additional discussion).
The description of the velocity of a point of a rotating rigid body is an important
example of the geometric and physical application of vectors. Suppose a rigid
body is rotating with an angular velocity ω about an axis and we wish to describe
the velocity of some point P of the body, as shown in Figure 2.3.8(a).
Let v denote the velocity at the point P. Each point of the body describes a
circle that lies in a plane perpendicular to the axis with its center on the axis. The
radius of the circle, a, is the perpendicular distance from the axis to the point
of interest. The magnitude of the velocity is equal to ωa. The direction of v is
perpendicular both to a and to the axis of rotation. We denote the direction of the
velocity by the unit vector ê. Thus, we can write

v = ω a ê. (2.3.14)

Let O be a reference point on the axis of revolution, and let OP = r. We then


have a = r sinθ, so that

v = ω r sin θ ê. (2.3.15)

The angular velocity is a vector because it has an assigned direction and magni-
tude, and obeys the parallelogram law of addition. We denote it by ω and represent
its direction in the sense of a right-handed screw, as shown in Figure 2.3.8(b). If
we further let êr be a unit vector in the direction of r, we see that
êω × êr = ê sin θ. (2.3.16)

ω
a êω
O θ ω
r P ê v
(a) (b)
(a) Velocity at a point in a rotating rigid body. (b) Angular velocity as a vector.
Figure 2.3.8
18 Principles of Continuum Mechanics

B×C A
C

B
Scalar triple product A · (B × C) as the volume of a parallelepiped.
Figure 2.3.9

With these relations, we have

v = ω × r. (2.3.17)

Thus, the velocity of a point of a rigid body rotating about an axis is given by the
vector product of ω and a position vector r drawn from any reference point on
the axis of revolution.
From the definition of the vector (cross) product, the following few simple
results follow:

(1) The products A × B and B × A are not equal. In fact, we have

A × B ≡ −B × A. (2.3.18)

Thus, the vector product does not commute. We must therefore preserve
the order of the vectors when vector products are involved.
(2) If two vectors A and B are parallel to each other, then θ = π or 0 and
sin θ = 0. Thus,

A × B = 0.

Conversely, if A × B = 0, then either A or B is zero or they are parallel


vectors. It follows that the vector product of a vector with itself is zero,
that is, A × A = 0.
(3) The distributive law still holds but the order of the factors must be main-
tained:

(A + B) × C = (A × C) + (B × C). (2.3.19)

2.3.7 Triple products of vectors


Now consider the various products of three vectors:
A(B · C), A · (B × C), A × (B × C). (2.3.20)

The product A(B · C) is merely a multiplication of the vector A by the scalar


B · C. On the other hand, the product A · (B × C) is a scalar and it is termed the
scalar triple product. It can be seen that the product A · (B × C), except for the
algebraic sign, is the volume of the parallelepiped formed by the vectors A, B,
and C, as shown in Figure 2.3.9.
19 Vectors and Tensors

B×C A C

n1C B
m1B A × (B × C), perpendicular to both A and B × C

The vector triple product.


Figure 2.3.10

We also note the following properties:

(1) The dot and cross operations can be interchanged without changing the
value:

A · B × C = A × B · C ≡ [ABC]. (2.3.21)

(2) A cyclical permutation of the order of the vectors leaves the result un-
changed:
A · B × C = C · A × B = B · C × A ≡ [ABC]. (2.3.22)

(3) If the cyclic order is changed, the sign changes:

A · B × C = −A · C × B = −C · B × A = −B · A × C. (2.3.23)

(4) A necessary and sufficient condition for any three vectors, A,B,C, to be
coplanar is that A · (B × C) = 0. Note also that the scalar triple product is
zero when any two vectors are the same.

The vector triple product A × (B × C) is a vector normal to the plane formed


by A and (B × C). However, the vector (B × C) is perpendicular to the plane
formed by B and C. This means that A × (B × C) lies in the plane formed by B
and C and is perpendicular to A, as shown in Figure 2.3.10. Thus, A × (B × C)
can be expressed as a linear combination of B and C:
A × (B × C) = c1 B + c2 C. (2.3.24)

Likewise, we would find that


(A × B) × C = d1 A + d2 B. (2.3.25)

Thus, the parentheses cannot be interchanged or removed. It can be shown that

c1 = A · C, c2 = −A · B,

and hence that


A × (B × C) = (A · C)B − (A · B)C. (2.3.26)
20 Principles of Continuum Mechanics

S Snˆ


B
C=A B
ê
S
A
(a) (b)
(a) Plane area as a vector. (b) Unit normal vector and sense of travel.
Figure 2.3.11

Example 2.3.1:
Let A and B be any two vectors in space. Express vector A in terms of its
components along (that is, parallel) and perpendicular to vector B.

Solution: The component of A along B is given by (A · ê B ), where ê B = B/B


is the unit vector in the direction of B. The component of A perpendicular to B
and in the plane of A and B is given by the vector triple product ê B × (A × ê B ).
Thus,

A = (A · ê B )ê B + ê B × (A × ê B ).

Alternatively, using Eq. (2.3.26) with A = C = ê B and B = A, we obtain


ê B × (A × ê B ) = A − (ê B · A)ê B

or

A = (A · ê B )ê B + ê B × (A × ê B ). (2.3.27)

2.3.8 Plane area as a vector


As indicated previously, the magnitude of the vector C = A × B is equal to
the area of the parallelogram formed by the vectors A and B, as shown in
Figure 2.3.11(a). In fact, the vector C may be considered to represent both the
magnitude and the direction of the product A and B. Thus, a plane area may be
looked upon as possessing a direction in addition to a magnitude, the directional
character rising out of the need to specify an orientation of the plane in space.
It is customary to denote the direction of a plane area by means of a unit vector
drawn normal to that plane. To fix the direction of the normal, we assign a sense
of travel along the contour of the boundary of the plane area in question. The
direction of the normal is taken by convention as that in which a right-handed
screw advances as it is rotated according to the sense of travel along the boundary
curve or contour, as shown in Figure 2.3.11(b). Let the unit normal vector be
given by n̂; then the area can be denoted by S = S n̂.
The representation of a plane as a vector has many uses. The vector can be
used to determine the area of an inclined plane in terms of its projected area, as
illustrated in the next example.
21 Vectors and Tensors

Example 2.3.2:

(1) Determine the plane area of the surface obtained by cutting a cylinder of
cross-sectional area S0 with an inclined plane whose normal is n̂, as shown
in Figure 2.3.12(a).
(2) Express the areas of the sides of the tetrahedron obtained from a cube
(or a prism) cut by an inclined plane whose normal is n̂, as shown in
Figure 2.3.12(b), in terms of the area S of the inclined surface.

Solution:
(1) Let the plane area of the inclined surface be S, as shown in Figure 2.3.12(a).
First, we express the areas as vectors,

S0 = S0 n̂0 and S = S n̂.

Because S0 is the projection of S along n̂0 (if the angle between n̂ and n̂0
is acute; otherwise the negative of it),

S0 = S · n̂0 = S n̂ · n̂0 . (2.3.28)

The scalar product n̂ · n̂0 is the cosine of the angle between the two unit
normal vectors.
(2) For reference purposes, we label the sides of the cube by 1, 2, and 3 and the
normals and surface areas by (n̂1 , S1 ), (n̂2 , S2 ), and(n̂3 , S3 ), respectively
(i.e., Si is the surface area of the plane perpendicular to the ith line or n̂i
vector), as shown in Figure 2.3.12(b). Then we have

n̂1 = −ê1 , n̂2 = −ê2 , n̂3 = −ê3 (2.3.29)

S1 = S n̂ · ê1 = Sn 1 , S2 = S n̂ · ê2 = Sn 2 , S3 = S n̂ · ê3 = Sn 3


(2.3.30)

x3


n̂1 x2

S S1 S

n̂0
n̂ 2
S0 S2
S3
n̂ 3 x1
(a) (b)
Vector representation of an inclined plane area.
Figure 2.3.12
22 Principles of Continuum Mechanics

A3e 3
A3ê 3
e3 A2 e 2
A ê3 A
A 1e1 A 1ê1 A2ê 2
e2
e1 ê1 ê2

(a) (b)
Components of a vector in (a) the general coordinate system that is oblique, and (b) the rectangular Cartesian
Figure 2.3.13
system.

2.3.9 Components of a vector


So far, we have considered a geometric description of a vector as a directed line
segment. We now embark on an analytic description of a vector and some of the
operations associated with this description. The analytic description of vectors
is useful in expressing, for example, the laws of physics in analytic form. The
analytic description of a vector is based on the notion of its components.
In a three-dimensional space, a set of no more than three linearly independent
vectors can be found. Let us choose any set and denote it as follows:

e1 , e 2 , e 3 . (2.3.31)

This set is called a basis (or a base system). A basis is called orthonormal if
they are mutually orthogonal and have unit magnitudes. To distinguish the basis
(e1 , e2 , e3 ) that is not orthonormal from one that is orthonormal, we denote the
orthonormal basis by (ê1 , ê2 , ê3 ), where
ê1 · ê2 = 0, ê2 · ê3 = 0, ê3 · ê1 = 0,
(2.3.32)
ê1 · ê1 = 1, ê2 · ê2 = 1, ê3 · ê3 = 1.

In some books, the notation (î, ĵ, k̂) or (êx , ê y , êz ) is used in place of (ê1 , ê2 , ê3 ).
In view of the previous discussion of the cross product of vectors, we note the
following relations resulting from the cross products of the basis vectors:
ê1 × ê1 = 0, ê1 × ê2 = ê3 , ê1 × ê3 = −ê2 ,
ê2 × ê1 = −ê3 , ê2 × ê2 = 0, ê2 × ê3 = ê1 , (2.3.33)
ê3 × ê1 = ê2 , ê3 × ê2 = −ê1 , ê3 × ê3 = 0.

It is clear from the concept of linear dependence that we can represent any
vector in three-dimensional space as a linear combination of the basis vectors:
A = A x êx + A y ê y + A z êz = A1 ê1 + A2 ê2 + A3 ê3 . (2.3.34)
The vectors A1 ê1 , A2 ê2 , and A3 ê3 are called the vector components of A, and
A1 , A2 , and A3 are called scalar components of A associated with the basis
(ê1 , ê2 , ê3 ), as indicated in Figure 2.3.13.
When the basis is orthonormal, A1 , A2 , and A3 are the physical components
of the vector A, that is, the components have the same physical dimensions or
23 Vectors and Tensors

units as the vector. A scalar multiple of a vector is the same as the vector whose
components are the scalar multiples:

αA = (α A1 )ê1 + (α A2 )ê2 + (α A3 )ê3 . (2.3.35)

Two vectors are equal if and only if their respective components are equal. That
is, A = B implies that A1 = B1 , A2 = B2 , and A3 = B3 .
The operations of vector addition, scalar product, and vector product of vectors
can now be expressed in terms of the rectangular Cartesian components, as given
in the following.
Addition of Vectors. The sum of vectors A and B is the vector C whose components
are the sum of the respective components of vectors A and B:
   
A + B = A1 ê1 + A2 ê2 + A3 ê3 + B1 ê1 + B2 ê2 + B3 ê3
= (A1 + B1 )ê1 + (A2 + B2 )ê2 + (A3 + B3 )ê3
≡ C1 ê1 + C2 ê2 + C3 ê3 = C, (2.3.36)

with C1 = A1 + B1 , C2 = A2 + B2 , and C3 = A3 + B3 .
Scalar Product of Vectors. The scalar product of vectors A and B is the scalar
   
A · B = A1 ê1 + A2 ê2 + A3 ê3 · B1 ê1 + B2 ê2 + B3 ê3
= A1 B1 + A2 B2 + A3 B3 , (2.3.37)

where the orthonormal property, Eq. (2.3.32), of the basis vectors is used in
arriving at this last expression.
Vector Product of Vectors. The vector product of two vectors A and B is the
vector
   
A × B = A1 ê1 + A2 ê2 + A3 ê3 × B1 ê1 + B2 ê2 + B3 ê3
= (A2 B3 − A3 B2 )ê1 + (A3 B1 − A1 B3 )ê2 + (A1 B2 − A2 B1 )ê3 , (2.3.38)

where the relations in Eq. (2.3.33) are used in arriving at the final expression.

Example 2.3.3:
The velocity at a point in a flow field is v = 2î + 3ĵ (m/s). Determine (1) the
velocity vector vn normal to the plane n = 3î − 4k̂ passing through the point,
(2) the angle between v and vn , (3) the tangential velocity vector vt on the
plane, and (4) the mass flow rate across the plane through an area A = 0.15
m2 if the density of the fluid (water) is ρ = 103 kg/m3 and the flow is uniform.
Solution: (1) The magnitude of the velocity normal to the given plane is given
by the projection of the velocity along the normal to the plane. The unit vector
normal to the plane is given by
1 
n̂ = 3î − 4k̂ .
5
24 Principles of Continuum Mechanics

A = 0.15m2
ˆ ˆ
v = 2i+3j

y vt
v
θ

ˆi
vn
x
kˆ n = 3iˆ - 4k
ˆ

z
(a) (b)
Flow across a plane.
Figure 2.3.14

Then the velocity vector normal to the plane is [see Figure 2.3.14(a)]
6  
n̂ = 0.24 3î − 4k̂ m/s.
vn = (v · n̂)n̂ =
5
(2) The angle between v and vn is given by
     
−1 v · vn −1 vn −1 1.2
θ = cos = cos = cos √ = 70.6◦ .
|v||vn | v 13
(3) The tangential velocity vector on the plane is given by
 
vt = v − vn = 2î + 3ĵ − 0.24 3î − 4k̂ = −5.2î + 3ĵ + 9.6k̂.

(4) The mass flow rate is given by

Q = ρvn A = 103 × 1.2 × 0.15 = 180 kg/s.


Various vectors are depicted in Figure 2.3.14(b).

2.4 Index notation and summation convention


2.4.1 Summation convention
The use of index notation facilitates writing long expressions in a succinct form.
For example, consider the component form of vector A,
A = A1 e1 + A2 e2 + A3 e3 , (2.4.1)

which can be abbreviated as



3 
3
A= Ai ei or A = Ajej.
i=1 j=1

If we had chosen the notation


A = A x e x + A y e y + A z ez ,

where (A x , A y , A z ) are the same components as ( A1 , A2 , A3 ) and the basis


(ex , e y , ez ) is the same as (e1 , e2 , e3 ), it is not possible to write this expression
25 Vectors and Tensors

with the summation convention. The summation index i or j is arbitrary as


long as the same index is used for both A and ê. The expression can be further
shortened by omitting the summation sign and having the understanding that a re-
peated index means summation over all values of that index. Thus, the three-term
expression A1 e1 + A2 e2 + A3 e3 can simply be written as
A = Ai ei . (2.4.2)
This notation is called the summation convention. The summation convention
allows us to write several expressions or equations in a single statement. As
we shall see shortly, the six relations in Eq. (2.3.32) can be written as a single
equation with the help of index notation and certain symbols that we are about to
introduce in Section 2.4.4.

2.4.2 Dummy index


The repeated index is called a dummy index because it can be replaced by any other
symbol that has not already been used in that expression. Thus, the expression in
Eq. (2.4.2) can also be written as
A = Ai ei = A j e j = Am em , (2.4.3)
and so on. As a rule, no index must appear more than twice in an expression. For
example, Ai Bi Ci is not a valid expression because the index i appears more than
twice. Other examples of dummy indices are
Fi = Ai B j C j , G k = Hk (2 − 3Ai Bi ) + P j Q j Fk .
Each of these equations expresses three equations when the range of i and j is
1 to 3. For example, the first equation is equal to the following three equations:
F1 = A1 (B1 C1 + B2 C2 + B3 C3 ),
F2 = A2 (B1 C1 + B2 C2 + B3 C3 ),
F3 = A3 (B1 C1 + B2 C2 + B3 C3 ).
This amply illustrates the usefulness of the summation convention in shortening
long and multiple expressions into a single expression.

2.4.3 Free index


A free index is one that appears in every expression of an equation except for
expressions that contain real numbers (scalars) only. The index i in the equation
Fi = Ai B j C j and k in the equation G k = Hk (2 − 3Ai Bi ) + P j Q j Fk are free
indices. Another example is
Ai = 2 + Bi + Ci + Di + (F j G j − H j P j )E i .
This expression contains three equations (i = 1, 2, 3). The expressions Ai =
B j Ck , Ai = B j , and Fk = Ai B j Ck do not make sense and should not arise
because the indices on the two sides of the equal sign do not match.
26 Principles of Continuum Mechanics

2.4.4 Kronecker delta and permutation symbols


It is convenient to introduce the Kronecker delta δi j and alternating symbol ei jk
because they allow easy representation of the dot product (or scalar product)
and cross product, respectively, of orthonormal vectors in a right-handed basis
system. We define the dot product êi · ê j as

êi · ê j = δi j , (2.4.4)

where

1, if i = j
δi j = (2.4.5)
0, if i = j.
Thus, the single expression in Eq. (2.4.4) is the same as the six relations in
Eq. (2.3.32). Due to its definition, the Kronecker delta δi j modifies (or contracts)
the subscripts in the coefficients of an expression in which it appears:

Ai δi j = A j , Ai B j δi j = Ai Bi = A j B j , δi j δik = δ jk .

As we shall see shortly, δi j denotes the Cartesian components of a second-order


unit tensor, I = δi j êi ê j = êi êi .
We define the cross product êi × ê j as
êi × ê j ≡ ei jk êk or ei jk = êi × ê j · êk = êi · ê j × êk , (2.4.6)

where


 1,if i, j, k are in cyclic order



 and not repeated (i = j = k),
ei jk = −1, if i, j, k are not in cyclic order (2.4.7)



 and not repeated (i = j = k),


0, if any of i, j, k are repeated.

The symbol ei jk is called the alternating symbol or permutation symbol. By


definition, the subscripts of the permutation symbol can be permuted without
changing its value; an interchange of any two subscripts will change the sign.
Hence, the interchange of two subscripts twice keeps the value unchanged:

ei jk = eki j = e jki , ei jk = −e jik = e jki = −ek ji .

In an orthonormal basis (e1 , e2 , e3 ), the scalar and vector products can be


expressed with the index notation using the Kronecker delta and the alternating
symbol:
A · B = (Ai êi ) · (B j ê j ) = Ai B j δi j = Ai Bi ,
(2.4.8)
A × B = (Ai êi ) × (B j ê j ) = Ai B j ei jk êk .
27 Vectors and Tensors

Note that the components of a vector in an orthonormal coordinate system can


be expressed as

Ai = A · êi , (2.4.9)

and therefore we can express vector A as

A = Ai êi = (A · êi )êi . (2.4.10)

Further, the Kronecker delta and the permutation symbol are related by an identity
known as the e-δ identity [see Problem 2.14],

ei jk eimn = δ jm δkn − δ jn δkm . (2.4.11)

The permutation symbol and the Kronecker delta prove to be very useful in
proving vector identities. Because a vector form of any identity is invariant (i.e.,
valid in any coordinate system), it suffices to prove it in only one coordinate
system. In particular, an orthonormal system is very convenient because we can
use the index notation, the permutation symbol, and the Kronecker delta. The
following examples contain several cases of incorrect and correct use of index
notation, and illustrate some of the uses of δi j and ei jk .

Example 2.4.1:
Discuss the validity of the following expressions:

(1) am bs = cm (dr − fr )
(2) am bs = cm (ds − f s )
(3) ai = b j ci di
(4) xi xi = r 2
(5) ai = 3

Solution:
(1) This is not a valid expression because the free indices r and s do not match.
(2) Valid; both m and s are free indices. There are nine equations (m, s =
1, 2, 3).
(3) This is not a valid expression because the free index j is not matched on
both sides of the equality, and index i is a dummy index in one expression
and a free index in the other. The index i cannot be used both as a free
and dummy index in the same equation. The equation would be valid if i
on the left side of the equation is replaced with j; then there will be three
equations).
(4) A valid expression, containing one equation: x 12 + x22 + x32 = r 2 .
(5) This is a valid expression in some branches of mathematics but it is
not a valid expression in continuum mechanics because it violates form-
invariance (material frame indifference) under a basis transformation
(every component of a vector cannot be the same in all bases).
28 Principles of Continuum Mechanics

Example 2.4.2:
Simplify the following expressions:

(1) δi j δ jk δkp δ pi
(2) εm jk εn jk
(3) (A × B) · (C × D)

Solution:
(1) Successive contraction of subscripts yield the result of

δi j δ jk δkp δ pi = δi j δ jk δki = δi j δ ji = δii = 3.

(2) Expand this expression using the e-δ identity:

εm jk εn jk = δmn δ j j − δm j δn j = 3δmn − δmn = 2δmn .


In particular, the expression εi jk εi jk is equivalent to 2δii = 6.
(3) Expanding the expression using index notation, we obtain

(A × B) · (C × D) = (Ai B j ei jk êk ) · (Cm Dn emnp ê p )

= Ai B j Cm Dn ei jk emnp δkp

= Ai B j Cm Dn ei jk emnk

= Ai B j Cm Dn (δim δ jn − δin δ jm )

= Ai B j Cm Dn δim δ jn − Ai B j Cm Dn δin δ jm

= Ai B j Ci D j − Ai B j C j Di

= Ai Ci B j D j − Ai Di B j C j

= (A · C)(B · D) − (A · D)(B · C),

where we have used the e-δ identity, Eq. (2.4.11). Although the previ-
ous vector identity is established in an orthonormal coordinate system, it
holds in a general coordinate system. That is, the vector identity here is
invariant.

Example 2.4.3:
Rewrite the expression emni Ai B j Cm Dn ê j in vector form.
Solution: We note that B j ê j = B. Examining the indices in the permutation
symbol and the remaining coefficients, it is clear that vectors C and D must
have a cross product between them and the resulting vector must have a dot
product with vector A. Thus, we have
emni Ai B j Cm Dn ê j = [(C × D) · A]B = (C × D · A) B.
29 Vectors and Tensors

z = x3

. ( x, y , z ) = ( x , x , x )
1 2 3
r
ê3, êz x3
ê1 , ê x
y = x2
êy , ê2
x1
x2
x = x1
Rectangular Cartesian coordinates.
Figure 2.4.1

2.4.5 Transformation law for different bases


When the basis vectors are constant, that is, with fixed lengths (with the same
units) and directions, the basis is called Cartesian. The general Cartesian system
is oblique. When the basis vectors are unit and orthogonal (orthonormal), the
basis system is called rectangular Cartesian, or simply Cartesian. In much of
our study, we shall deal with Cartesian bases.
Let us denote an orthonormal Cartesian basis by

{êx , ê y , êz } or {ê1 , ê2 , ê3 }.

The Cartesian coordinates are denoted by (x, y, z) or (x 1 , x2 , x3 ). The familiar


rectangular Cartesian coordinate system is shown in Figure 2.4.1. We shall always
use right-handed coordinate systems.
A position vector to an arbitrary point (x, y, z) or (x 1 , x2 , x3 ), measured from
the origin, is given by

r = x êx + y ê y + z êz
= x1 ê1 + x2 ê2 + x3 ê3 , (2.4.12)

or in summation notation by

r = x j ê j , r · r = r 2 = xi xi . (2.4.13)

We shall also use the symbol x for the position vector r = x. The length of a line
element dr = dx is given by
dr · dr = (ds)2 = d x j d x j = (d x)2 + (dy)2 + (dz)2 . (2.4.14)

Here we discuss the relationship between the components of two different


orthonormal coordinate systems. Consider the first coordinate basis

{ê1 , ê2 , ê3 }


and the second coordinate basis

{ēˆ 1 , ēˆ 2 , ēˆ 3 }.


30 Principles of Continuum Mechanics

Now we can express the same vector in the coordinate system without bars
(referred to as “unbarred”), and also in the coordinate system with bars (referred
to as “barred”):
A = Ai êi = (A · êi )êi
(2.4.15)
= Āi ēˆ i = (A · ēˆ i )ēˆ i .
From Eq. (2.4.10), we have
Ā j = A · ēˆ j = Ai (êi · ēˆ j ) ≡ ji Ai , (2.4.16)
where

ij = ēˆ i · ê j . (2.4.17)
Equation (2.4.12) gives the relationship between the components ( Ā1 , Ā2 , Ā3 )
and (A1 , A2 , A3 ), and this is called the transformation rule between the barred
and unbarred components in the two coordinate systems. The coefficients i j can
be interpreted as the directional cosines of the barred coordinate system with
respect to the unbarred coordinate system:

ij = cosine of the angle between ēˆ i and ê j . (2.4.18)


Note that the first subscript of i j comes from the barred coordinate system and the
second subscript from the unbarred system. Obviously, i j is not symmetric (i.e.,
i j = ji ). The rectangular array of these components is called a matrix, which
is the topic of the next section. The next example illustrates the computation of
directional cosines.

Example 2.4.4:
Let êi (i = 1, 2, 3) be a set of orthonormal base vectors, and define a new
right-handed coordinate basis by (note that ēˆ 1 · ēˆ 2 = 0)
1 1
ēˆ 1 = (2ê1 + 2ê2 + ê3 ) , ēˆ 2 = √ (ê1 − ê2 ) ,
3 2
1
ēˆ 3 = ēˆ 1 × ēˆ 2 = √ (ê1 + ê2 − 4ê3 ) .
3 2
The original and new coordinate systems are depicted in Figure 2.4.2. Deter-
mine the directional cosines i j of the transformation.
Solution: From Eq. (2.4.17), we have
2 2 1
11 = ēˆ 1 · ê1 = , 12 = ēˆ 1 · ê2 = , 13 = ēˆ 1 · ê3 = ,
3 3 3
1 1
21 = ēˆ 2 · ê1 = √ , 22 = ēˆ 2 · ê2 = − √ , 23 = ēˆ 2 · ê3 = 0,
2 2
1 1 4
31 = ēˆ 3 · ê1 = √ , 32 = ēˆ 3 · ê2 = √ , 33 = ēˆ 3 · ê3 = − √ .
3 2 3 2 3 2
31 Vectors and Tensors

ê3

ê2

ê2
eˆ1
ê1
ê3
The original and transformed coordinate systems defined in Example 2.4.4.
Figure 2.4.2

2.5 Theory of matrices


2.5.1 Definition
In the preceding sections, we studied the algebra of ordinary vectors and the
transformation of vector components from one coordinate system to another. For
example, the transformation equation Eq. (2.4.16) relates the components of a
vector in the barred coordinate system to an unbarred coordinate system. Writing
Eq. (2.4.16) in an expanded form,
Ā1 = 11 A1 + 12 A2 + 13 A3 ,

Ā2 = 21 A1 + 22 A2 + 23 A3 , (2.5.1)
Ā3 = 31 A1 + 32 A2 + 33 A3 ,

we see that there are nine coefficients relating the components Ai to Āi . The form
of these linear equations suggests writing the scalars of i j ( jth components in a
ith equation) in a rectangular array,
 
11 12 13
L= 21 22 23 .
31 32 33

This rectangular array L of scalars i j is called a matrix, and the quantities i j


are called the elements of L.2
If a matrix has m rows and n columns, we will say that is an m by n (m × n)
matrix, the number of rows always being listed first. The element in the ith row
and jth column of a matrix A is generally denoted by ai j , and we will sometimes
designate a matrix by A = [A] = [ai j ]. A square matrix is one that has the same
number of rows as columns, n × n. An n × n matrix is said to be of order n. The
elements of a square matrix for which the row number and the column number are
2 The word “matrix” was first used in 1850 by James Sylvester (1814–1897), an English algebraist.
However, Arthur Caley (1821–1895), professor of mathematics at Cambridge, was the first one
to explore the properties of matrices. Significant contributions in the early years were made by
Charles Hermite, Georg Frobenius, and Camille Jordan, among others.
32 Principles of Continuum Mechanics

the same (that is, aii for any fixed i) are called diagonal elements, or simply the
diagonal. A square matrix is said to be a diagonal matrix if all of the off-diagonal
elements are zero. An identity matrix, denoted by I = [I ], is a diagonal matrix
whose elements are all 1’s. Examples of a diagonal and an identity matrix are
given as


5 0 0 0 
1 0 0 0
0 −2 0 0 0 1 0 0
A=
0
 and I= .
0 1 0 0 0 1 0
0 0 0 3 0 0 0 1
The sum of the diagonal elements is called the trace of the matrix.
If the matrix has only one row or one column, we will normally use only a
single subscript to designate its elements. For example,
 
 x1 
X = x2 , Y = {y1 y2 y3 }
 
x3
denote a column matrix and a row matrix, respectively. Row and column matrices
can be used to denote the components of a vector.

2.5.2 Matrix addition and multiplication of a matrix by a scalar


The sum of two matrices of the same size is defined to be a matrix of the same
size obtained by simply adding the corresponding elements. If A is an m × n
matrix and B is an m × n matrix, their sum is an m × n matrix, C, with
ci j = ai j + bi j for all i, j. (2.5.2)
A constant multiple of a matrix is equal to the matrix obtained by multiplying
all of the elements by the constant. That is, the multiple of a matrix A by a scalar
α, αA, is the matrix obtained by multiplying each of its elements by α:
 
a11 a12 . . . a1n  
 a21 a22 . . . a2n  αa11 αa12 . . . αa1n
 
A =  .. .. ..  , αA =  αa21 αa22 . . . αa2n  .
 . . ... .  αam1 αam2 . . . αamn
am1 am2 . . . amn
Matrix addition has the following properties:

(1) Addition is commutative: A + B = B + A.


(2) Addition is associative: A + (B + C) = (A + B) + C.
(3) There exists a unique matrix 0, such that A + 0 = 0 + A = A. The matrix
0 is called the zero matrix; with all elements of it are zeros.
(4) For each matrix A, there exists a unique matrix −A such that A +
(−A) = 0.
(5) Addition is distributive with respect to scalar multiplication: α(A + B) =
αA + αB.
33 Vectors and Tensors

(6) Addition is distributive with respect to matrix multiplication, which will


be discussed shortly (note the order):

(A + B)C = AC + BC.
Calculations of the sum and difference of matrices are illustrated through
the next example.

Example 2.5.1:
Compute the sum and difference of the following two matrices:

5 −2 12 21  
13 −11 32 4
 10 2 16 −3   −6 32 25 7 
A=
 20
, B= .
14 13 8  39 36 −23 15 
−12 31 0 19 14 −15 31 18
Solution: The sum of A and B is

5 −2 12 21   13 −11 32 4 
 10 −3   25 7 
A+B=
2 16  +  −6 32 
 20 14 13 8   39 36 −23 15 
−12 31 0 19 14 −15 31 18

18 −13 44 25 
 4 34 41 4
= 59
.
50 −10 23 
2 16 31 37
The difference of A and B is

5 −2 12 21   13 −11 32 4 
 10 16 −3   25 7 
A−B=
2  −  −6 32 
 20 14 13 8   39 36 −23 15 
−12 31 0 19 14 −15 31 18
 
−8 9 −20 17
 16 −30 −9 −10 
=  −19 −22
.
36 −7 
−26 46 −31 1

2.5.3 Matrix transpose and symmetric and skew symmetric matrices


If A is an m × n matrix, then the n × m matrix obtained by interchanging its rows
and columns is called the transpose of A and is denoted by AT . For example,
consider the matrices

5 −2 1   
3 −1 2 4
 8 7 6 
A=  2
 , B =  −6 3 5 7. (2.5.3)
4 3
9 6 −2 1
−1 9 0
34 Principles of Continuum Mechanics

The transpose matrices of A and B are



  3 −6 9
5 8 2 −1
 −1 3 6
AT =  −2 7 4 9  , BT = 
 2
.
5 −2 
1 6 3 0
4 7 1

The following basic properties of a transpose matrix should be noted:

(1) (AT )T = A.
(2) (A + B)T = AT + BT .

A square matrix A of real numbers is said to be symmetric if AT = A. It is said


to be skew symmetric or antisymmetric if AT = −A. In terms of the elements of
A, these definitions imply that A is symmetric if and only if ai j = a ji , and it is
skew symmetric if and only if ai j = −a ji . Note that the diagonal elements of a
skew symmetric matrix are always zero because ai j = −ai j implies ai j = 0 for
i = j. Examples of symmetric and skew symmetric matrices, respectively, are

5 −2 12 21  
0 −11 32 4 
 −2 2 16 −3   11 0 25 7 
 ,  .
 12 16 13 8  −32 −25 0 15 
21 −3 8 19 −4 −7 −15 0

2.5.4 Matrix multiplication


Consider a vector A = a1 ê1 + a2 ê2 + a3 ê3 in a Cartesian system. We can repre-
sent A as a product of a row matrix and a column matrix,
 
 ê1 
A = {a1 a2 a3 } ê2 .
 
ê3
Note that the vector A is obtained by multiplying the ith element in the row matrix
with the ith element in the column matrix, and then adding terms. This gives us
a strong motivation for defining the product of two matrices.
Let x and y be the vectors (matrices with one column)
   
 x1   y1 

 
 
 
 x2   y2  
x= .. , y= .. .

 . 
 
 . 


   
  
xm ym
We define the product xT y to be the scalar
 
 y1 

 
 y2  m
x y = {x1 , x2 , . . . , xm }
T
.. = x y
1 1 + x y
2 2 + · · · + x y
m m = xi yi .

 . 


   i=1
ym
(2.5.4)
Discovering Diverse Content Through
Random Scribd Documents
domestic life; he will blot out precise details, special traits, and will
carry tragedy into a serene and sublime region, where his abstract
personages, unencumbered by time and space, after an exchange of
eloquent harangues and able dissertations, will kill each other
becomingly, and as though they were merely concluding a ceremony.
Shakespeare does just the contrary, because his genius is the exact
opposite. His master faculty is an impassioned imagination, freed
from the shackles of reason and morality. He abandons himself to it,
and finds in man nothing that he would care to lop off. He accepts
nature and finds it beautiful in its entirety. He paints it in its
littlenesses, it deformities, its weaknesses, its excesses, its
irregularities, and its rages; he exhibits man at his meals, in bed, at
play, drunk, mad, sick; he adds that which ought not to be seen to
that which passes on the stage. He does not dream of ennobling,
but of copying human life, and aspires only to make his copy more
energetic and more striking than the original.
Hence the morals of this drama; and first, the want of dignity.
Dignity arises from self-command. A man selects the most noble of
his acts and attitudes, and allows himself no other. Shakespeare's
characters select none, but allow themselves all. His kings are men,
and fathers of families. The terrible Leontes, who is about to order
the death of his wife and his friend, plays like a child with his son:
caresses him, gives him all the pretty pet names which mothers are
wont to employ; he dares be trivial; he gabbles like a nurse; he has
her language and fulfils her duties:
"Leontes. What, hast smutch'd thy nose?
They say it is a copy out of mine. Come, captain,
We must be neat; not neat, but cleanly, captain:...
Come, sir page,
Look on me with your welkin eye: sweet villain!
Most dear'st! my collop... Looking on the lines
Of my boy's face, methoughts I did recoil
Twenty-three years, and saw myself unbreech'd,
In my green velvet coat, my dagger muzzled,
Lest it should bite its master....
How like, methought, I then was to this kernel,
This squash, this gentleman!... My brother,
Are you so fond of your young prince as we
Do seem to be of ours?
Polixenes. If at home, sir,
He's all my exercise, my mirth, my matter,
Now my sworn friend and then mine enemy,
My parasite, my soldier, statesman, all:
He makes a July's day short as December,
And with his varying childness cures in me
Thoughts that would thick my blood."[636]
There are a score of such passages in Shakespeare. The great
passions, with him as in nature, are preceded or followed by trivial
actions, small-talk, commonplace sentiments. Strong emotions are
accidents in our life: to drink, to eat, to talk of indifferent things, to
carry out mechanically a habitual duty, to dream of some stale
pleasure or some ordinary annoyance, that is in which we employ all
our time. Shakespeare paints us as we are; his heroes bow, ask
people for news, speak of rain and fine weather, as often and as
casually as ourselves, on the very eve of falling into the extremity of
misery, or of plunging into fatal resolutions. Hamlet asks what's
o'clock, finds the wind biting, talks of feasts and music heard
without; and this quiet talk, so unconnected with the action, so full
of slight, insignificant facts, which chance alone has raised up and
guided, lasts until the moment when his father's ghost, rising in the
darkness, reveals the assassination which it is his duty to avenge.
Reason tells us that our manners should be measured; this is why
the manners which Shakespeare paints are not so. Pure nature is
violent, passionate: it admits no excuses, suffers no middle course,
takes no count of circumstances, wills blindly, breaks out into railing,
has the irrationality, ardor, anger of children. Shakespeare's
characters have hot blood and a ready hand. They cannot restrain
themselves, they abandon themselves at once to their grief,
indignation, love, and plunge desperately down the steep slope,
where their passion urges them. How many need I quote? Timon,
Posthumus, Cressida, all the young girls, all the chief characters in
the great dramas; everywhere Shakespeare paints the unreflecting
impetuosity of the impulse of the moment. Capulet tells his daughter
Juliet that in three days she is to marry Earl Paris, and bids her be
proud of it; she answers that she is not proud of it, and yet she
thanks the earl for this proof of love. Compare Capulet's fury with
the anger of Orgon,[637] and you may measure the difference of the
two poets and the two civilizations:
"Capulet. How now, how now, chop-logic! What is this?
'Proud,' and 'I thank you,' and 'I thank you not;'
And yet 'not proud,' mistress minion, you,
Thank me no thankings, nor proud me no prouds,
But fettle your fine joints 'gainst Thursday next,
To go with Paris to Saint Peter's church,
Or I will drag thee on a hurdle thither.
Out, you green-sickness carrion! out, you baggage!
You tallow-face!
Juliet. Good father, I beseech you on my knees,
Hear me with patience but to speak a word.
C. Hang thee, young baggage! disobedient wretch
I tell thee what: get thee to church o' Thursday,
Or never after look me in the face:
Speak not, reply not, do not answer me;
My fingers itch....
Lady C. You are too hot.
C. God's bread! it makes me mad:
Day, night, hour, tide, time, work, play,
Alone, in company, still my care hath been
To have her match'd: and having now provided
A gentleman of noble parentage,
Of fair demesnes, youthful, and nobly train'd,
Stuff'd, as they say, with honorable parts,
Proportion'd as one's thoughts would wish a man;
And then to have a wretched puling fool,
A whining mammet, in her fortune's tender,
To answer, 'I'll not wed; I cannot love,
I am too young; I pray you, pardon me,'—
But, an you will not wed, I'll pardon you:
Graze where you will, you shall not house with me:
Look to't, think on't, I do not use to jest.
Thursday is near; lay hand on heart, advise:
An you be mine, I'll give you to my friend;
An you be not, hang, beg, starve, die in the streets,
For, by my soul, I'll ne'er acknowledge thee."[638]
This method of exhorting one's child to marry is peculiar to
Shakespeare and the sixteenth century. Contradiction to these men
was like a red rag to a bull; it drove them mad.
We might be sure that in this age, and on this stage, decency was a
thing unknown. It is wearisome, being a check; men got rid of it,
because it was wearisome. It is a gift of reason and morality; as
indecency is produced by nature and passion. Shakespeare's words
are too indecent to be translated. His characters call things by their
dirty names, and compel the thoughts to particular images of
physical love. The talk of gentlemen and ladies is full of coarse
allusions; we should have to find out an alehouse of the lowest
description to hear like words nowadays.[639]
It would be in an alehouse too that we should have to look for the
rude jests and brutal kind of wit which form the staple of these
conversations. Kindly politeness is the slow fruit of advanced
reflection; it is a sort of humanity and kindliness applied to small
acts and everyday discourse; it bids man soften towards others, and
forget himself for the sake of others; it constrains genuine nature,
which is selfish and gross. This is why it is absent from the manners
of the drama we are considering. You will see carmen, out of
sportiveness and good humor, deal one another hard blows; so it is
pretty well with the conversation of the lords and ladies of
Shakespeare who are in a sportive mood; for instance, Beatrice and
Benedick, very well bred folk as things go,[640] with a great
reputation for wit and politeness, whose smart retorts create
amusement for the bystanders. These "skirmishes of wit" consist in
telling one another plainly: You are a coward, a glutton, an idiot, a
buffoon, a rake, a brute! You are a parrot's tongue, a fool, a... (the
word is there). Benedick says:
"I will go... to the Antipodes... rather than hold three
words' conference with this harpy.... I cannot endure my
Lady Tongue....
Don Pedro. You have put him down, lady, you have put him down.
Beatrice. So I would not he should do me, my lord, lest I should
prove the mother of fools."[641]
We can infer the tone they use when in anger. Emilia, in "Othello,"
says:
"He call'd her whore; a beggar in his drink
Could not have laid such terms upon his callat."[642]
They have a vocabulary of foul words as complete as that of
Rabelais, and they exhaust it. They catch up handfuls of mud and
hurl it at their enemy, not conceiving themselves to be smirched.
Their actions correspond. They go without shame or pity to the
limits of their passion. They kill, poison, violate, burn; the stage is
full of abominations. Shakespeare lugs upon the stage all the
atrocious deeds of the Civil Wars. These are the ways of wolves and
hyenas. We must read of Jack Cade's sedition[643] to gain an idea of
this madness and fury. We might imagine we were seeing infuriated
beasts, the murderous recklessness of a wolf in a sheepfold, the
brutality of a hog fouling and rolling himself in filth and blood. They
destroy, kill, butcher each other; with their feet in the blood of their
victims, they call for food and drink; they stick heads on pikes and
make them kiss one another, and they laugh.

"Jack Cade. There shall be in England seven halfpenny loaves


sold for a penny.... There shall be no money; all shall eat and
drink on my score, and I will apparel them all in one livery....
And here sitting upon London-stone, I charge and command
that, of the city's cost, the pissing-conduit run nothing but claret
wine this first year of our reign.... Away, burn all the records of
the realm; my mouth shall be the parliament of England.... And
henceforth all things shall be in common.... What canst thou
answer to my majesty for giving up of Normandy unto
Mounsieur Basimecu, the dauphin of France?... The proudest
peer in the realm shall not wear a head on his shoulders, unless
he pay me tribute; there shall not a maid be married, but she
shall pay to me her maidenhead ere they have it. (Re-enter
rebels with the heads of Lord Say and his son-in-law.) But is not
this braver? Let them kiss one another, for they loved well when
they were alive."[644]

Man must not be let loose; we know not what lusts and rage may
brood under a sober guise. Nature was never so hideous, and this
hideousness is the truth.
re these cannibal manners only met with among the scum? Why, the
princes are worse. The Duke of Cornwall orders the old Earl of
Gloucester to be tied to a chair, because, owing to him, King Lear
has escaped:
"Fellows, hold the chair.
Upon these eyes of thine I'll set my foot.
(Gloucester is held down in the chair, while Cornwall
plucks
out one of his eyes, and sets his foot on it.)
Glou. He that will think to live till he be old,
Give me some help! O cruel: O you gods!
Regan. One side will mock another; the other too.
Cornwall. If you see vengeance—
Servant. Hold your hand, my lord:
I have served you ever since I was a child;
But better service have I never done you,
Than now to bid you hold.
Regan. How now, you dog!
Serv. If you did wear a beard upon your chin,
I'd shake it on this quarrel. What do you mean?
Corn. My villain! (Draws and runs at him.)
Serv. Nay, then, come on, and take the chance of
anger.
(Draws; they fight; Cornwall is wounded.)
Regan. Give me thy sword. A peasant stands up thus.
(Snatches a sword, comes behind, and stabs him.)
Serv. O, I am slain! My lord, you have one eye left
To see some mischief on him. O! (Dies.)
Corn. Lest it see more, prevent it. Out, vile jelly!
Where is thy lustre now?
Glou. All dark and comfortless. Where's my son?...
Regan. Go thrust him out at gates, and let him smell
His way to Dover."[645]
Such are the manners of that stage. They are unbridled, like those
of the age, and like the poet's imagination. To copy the common
actions of every-day life, the puerilities and feeblenesses to which
the greatest continually sink, the outbursts of passion which degrade
them, the indecent, harsh, or foul words, the atrocious deeds in
which license revels, the brutality and ferocity of primitive nature, is
the work of a free and unencumbered imagination. To copy this
hideousness and these excesses with a selection of such familiar,
significant, precise details, that they reveal under every word of
every personage a complete civilization, is the work of a
concentrated and all-powerful imagination. This species of manners
and this energy of description indicate the same faculty, unique and
excessive, which the style had already indicated.

SECTION IV.—Dramatis Personæ

On this common background stands out in striking relief a population


of distinct living figures, illuminated by an intense light. This creative
power is Shakespeare's great gift, and it communicates an
extraordinary significance to his words. Every phrase pronounced by
one of its characters enables us to see, besides the idea which it
contains and the emotion which prompted it, the aggregate of the
qualities and the entire character which produced it—the mood,
physical attitude, bearing, look of the man, all instantaneously, with
a clearness and force approached by no one. The words which strike
our ears are not the thousandth part of those we hear within; they
are like sparks thrown off here and there; the eyes catch rare flashes
of flame; the mind alone perceives the vast conflagration of which
they are the signs and the effect. He gives us two dramas in one:
the first strange, convulsive, curtailed, visible; the other consistent,
immense, invisible; the one covers the other so well, that as a rule
we do not realize that we are perusing words: we hear the roll of
those terrible voices, we see contracted features, glowing eyes,
pallid faces; we see the agitation, the furious resolutions which
mount to the brain with the feverish blood, and descend to the
sharp-strung nerves. This property possessed by every phrase to
exhibit a world of sentiments and forms, comes from the fact that
the phrase is actually caused by a world of emotions and images.
Shakespeare, when he wrote, felt all that we feel, and much besides.
He had the prodigious faculty of seeing in a twinkling of the eye a
complete character, body, mind, past and present, in every detail
and every depth of his being, with the exact attitude and the
expression of face, which the situation demanded. A word here and
there of Hamlet or Othello would need for its explanation three
pages of commentaries; each of the half-understood thoughts, which
the commentator may have discovered, has left its trace in the turn
of the phrase, in the nature of the metaphor, in the order of the
words; nowadays, in pursuing these traces, we divine the thoughts.
These innumerable traces have been impressed in a second, within
the compass of a line. In the next line there are as many, impressed
just as quickly, and in the same compass. You can gauge the
concentration and the velocity of the imagination which creates thus.
These characters are all of the same family. Good or bad, gross or
delicate, witty or stupid, Shakespeare gives them all the same kind
of spirit which is his own. He has made of them imaginative people,
void of will and reason, impassioned machines, vehemently jostled
one against another, who were outwardly whatever is most natural
and most abandoned in human nature. Let us act the play to
ourselves, and see in all its stages this clanship of figures, this
prominence of portraits.
Lowest of all are the stupid folk, babbling or brutish. Imagination
already exists there, where reason is not yet born; it exists also
there where reason is dead. The idiot and the brute blindly follow
the phantoms which exist in their benumbed or mechanical brains.
No poet has understood this mechanism like Shakespeare. His
Caliban, for instance, a deformed savage, fed on roots, growls like a
beast under the hand of Prospero, who has subdued him. He howls
continually against his master, though he knows that every curse will
be paid back with "cramps and aches." He is a chained wolf,
trembling and fierce, who tries to bite when approached, and who
crouches when he see's the lash raised. He has a foul sensuality, a
loud base laugh, the gluttony of degraded humanity. He wishes to
violate Miranda in her sleep. He cries for his food, and gorges
himself when he gets it. A sailor who had landed in the island,
Stephano, gives him wine; he kisses his feet, and takes him for a
god; he asks if he has not dropped from heaven, and adores him.
We find in him rebellious and baffled passions, which are eager to
rise again and to be satiated. Stephano had beaten his comrade.
Caliban cries, "Beat Him enough: after a little time I'll beat him too."
He prays Stephano to come with him and murder Prospero in his
sleep; he thirsts to lead him there, dances through joy and sees his
master already with his "weasand" cut, and his brains scattered on
the earth:
"Prithee, my king, be quiet. See'st thou here,
This is the mouth o' the cell: no noise, and enter.
Do that good mischief which may make this island
Thine own forever, and I, thy Caliban,
For aye thy foot-licker."[646]
Others, like Ajax and Cloten, are more like men, and yet it is pure
mood that Shakespeare depicts in them, as in Caliban. The clogging
corporeal machine, the mass of muscles, the thick blood sluggishly
moving along in the veins of these fighting men, oppress the
intelligence, and leave no life but for animal passions. Ajax uses his
fists, and devours meat; that is his existence; if he is jealous of
Achilles, it is pretty much as a bull is jealous of his fellow. He permits
himself to be restrained and led by Ulysses, without looking before
him: the grossest flattery decoys him. The Greeks have urged him to
accept Hector's challenge. Behold him puffed up with pride, scorning
to answer anyone, not knowing what he says or does. Thersites
cries, "Good-morrow, Ajax"; and he replies, "Thanks, Agamemnon."
He has no further thought than to contemplate his enormous frame,
and roll majestically his big stupid eyes. When the day of the fight
has come, he strikes at Hector as on an anvil. After a good while
they are separated. "I am not warm yet," says Ajax, "let us fight
again."[647] Cloten is less massive than this phlegmatic ox; but he is
just as idiotic, just as vainglorious, just as coarse. The beautiful
Imogen, urged by his insults and his scullion manners, tells him that
his whole body is not worth as much a Posthumus's meanest
garment. He is stung to the quick, repeats the words several times;
he cannot shake off the idea, and runs at it again and again with his
head down, like an angry ram:
"Cloten. 'His garment?' Now, the devil—
Imogen. To Dorothy my woman hie thee presently—
C. 'His garment?'... You have abused me: 'His meanest
garment!'... I'll be revenged: 'His meanest garment!'
Well."[648]
He gets some of Posthumus's garments, and goes to Milford Haven,
expecting to meet Imogen there. On his way he mutters thus:

"With that suit upon my back, will I ravish her: first kill him, and
in her eyes; there shall she see my valor, which will then be a
torment to her contempt. He on the ground, my speech of
insultment ended on his dead body, and when my lust has dined
—which, as I say, to vex her I will execute in the clothes that
she so praised—to the court I'll knock her back, foot her home
again."[649]

Others again, are but babblers: for example, Polonius, the grave
brainless counsellor; a great baby, not yet out of his "swathing
clouts"; a solemn booby, who rains on men a shower of counsels,
compliments, and maxims; a sort of court speaking-trumpet, useful
in grand ceremonies, with the air of a thinker, but fit only to spout
words. But the most complete of all these characters is that of the
nurse in "Romeo and Juliet," a gossip, loose in her talk, a regular
kitchen oracle, smelling of the stewpan and old boots, foolish,
impudent, immoral, but otherwise a good creature, and affectionate
to her nurse-child. Mark this disjointed and never-ending gossip's
babble:
"Nurse. 'Faith I can tell her age unto an hour.
Lady Capulet. She's not fourteen....
Nurse. Come Lammas-eve at night shall she be
fourteen.
Susan and she—God rest all Christian souls!—
Were of an age: well, Susan is with God;
She was too good for me: but, as I said,
On Lammas-eve at night shall she be fourteen;
That shall she, marry; I remember it well.
'Tis since the earthquake now eleven years;
And she was wean'd—I never shall forget it—
Of all the days of the year, upon that day:
For I had then laid wormwood to my dug,
Sitting in the sun under the dove-house wall;
My lord and you were then at Mantua:—
Nay, I do bear a brain:—but, as I said,
When it did taste the wormwood on the nipple
Of my dug and felt it bitter, pretty fool,
To see it tetchy and fall out with the dug!
Shake, quoth the dove-house: 'twas no need, I trow,
To bid me trudge:
And since that time it is eleven years;
For then she could stand alone; nay, by the rood,
She could have run and waddled all about;
For even the day before, she broke her brow."[650]
Then she tells an indecent anecdote, which she begins over again
four times. She is silenced: what then? She has her anecdote in her
head, and cannot cease repeating it and laughing to herself. Endless
repetitions are the mind's first step. The vulgar do not pursue the
straight line of reasoning and of the story; they repeat their steps,
as it were merely marking time: struck with an image, they keep it
for an hour before their eyes, and are never tired of it. If they do
advance, they turn aside to a hundred subordinate ideas before they
get at the phrase required. They allow themselves to be diverted by
all the thoughts which come across them. This is what the nurse
does; and when she brings Juliet news of her lover, she torments
and wearies her, less from a wish to tease than from a habit of
wandering from the point:
"Nurse. Jesu, what haste? can you not stay awhile?
Do you not see that I am out of breath?
Juliet. How art thou out of breath, when thou hast
breath
To say to me that thou art out of breath?
Is thy news good, or bad? answer to that;
Say either, and I'll stay the circumstance:
Let me be satisfied: is't good or bad?
N. Well, you have made a simple choice; you know not
how to choose
a man: Romeo! no, not he: though his face be better
than any man's,
yet his legs excels all men's; and for a hand and a foot,
and a body,
though they be not to be talked on, yet they are past
compare: he is
not the flower of courtesy, but, I'll warrant him, as gentle
as a lamb.
Go thy ways, wench; serve God. What, have you dined at
home?
J. No, no: but all this did I know before.
What says he of our marriage? what of that?
N. Lord, how my head aches! what a head have I!
It beats as it would fall in twenty pieces.
My back o' t'other side—O, my back, my back!
Beshrew your heart for sending me about,
To catch my death with jaunting up and down!
J. I' faith, I am sorry that thou art not well.
Sweet, sweet, sweet nurse, tell me, what says my love?
N. Your love says, like an honest gentleman, and a
courteous, and
a kind, and a handsome, and, I warrant, a virtuous—
Where is your
mother?"[651]
It is never-ending. Her gabble is worse when she comes to
announce to Juliet the death of her cousin and the banishment of
Romeo. It is the shrill cry and chatter of an overgrown asthmatic
magpie. She laments, confuses the names, spins roundabout
sentences, ends by asking for aqua-vitœ. She curses Romeo, then
brings him to Juliet's chamber. Next day Juliet is ordered to marry
Earl Paris; Juliet throws herself into her nurse's arms, praying for
comfort, advice, assistance. The other finds the true remedy: Marry
Paris,
"O, he's a lovely gentleman!
Romeo's a dishclout to him: an eagle, madam,
Hath not so green, so quick, so fair an eye
As Paris hath. Beshrew my very heart,
I think you are happy in this second match.
For it excels your first."[652]
This cool immorality, these weather-cock arguments, this fashion of
estimating love like a fishwoman, completes the portrait.

SECTION V.—Men of Wit

The mechanical imagination produces Shakespeare's fool-characters:


a quick, venturesome, dazzling, unquiet imagination, produces his
men of wit. Of wit there are many kinds. One, altogether French,
which is but reason, a foe to paradox, scorner of folly, a sort of
incisive common-sense, having no occupation but to render truth
amusing and evident, the most effective weapon with an intelligent
and vain people: such was the wit of Voltaire and the drawing-
rooms. The other, that of improvisators and artists, is a mere
inventive rapture, paradoxical, unshackled, exuberant, a sort of self-
entertainment, a phantasmagoria of images, flashes of wit, strange
ideas, dazing and intoxicating, like the movement and illumination in
a ball-room. Such is the wit of Mercutio, of the clowns, of Beatrice,
Rosalind, and Benedick. They laugh, not from a sense of the
ridiculous, but from the desire to laugh. You must look elsewhere for
the campaigns with aggressive reason makes against human folly.
Here folly is in its full bloom. Our folk think of amusement, and
nothing more. They are good-humored; they let their wit prance
gayly over the possible and the impossible. They play upon words,
contort their sense, draw absurd and laughable inferences, send
them back to one another, and without intermission, as if with
shuttlecocks, and vie with each other in singularity and invention.
They dress all their ideas in strange or sparkling metaphors. The
taste of the time was for masquerades; their conversation is a
masquerade of ideas. They say nothing in a simple style; they only
seek to heap together subtle things, far-fetched, difficult to invent
and to understand; all their expressions are over-refined,
unexpected, extraordinary; they strain their thought, and change it
into a caricature. "Alas, poor Romeo!" says Mercutio, "he is already
dead; stabbed with a white wench's black eye; shot through the ear
with a love-song, the very pin of his heart cleft with the blind bow-
boy's butt-shaft."[653] Benedick relates a conversation he has just
held with his mistress: "O, she misused me past the endurance of a
block! an oak, but with one green leaf on it would have answered
her; my very visor began to assume life, and scold with her."[654]
These gay and perpetual extravagances show the bearing of the
speakers. They do not remain quietly seated in their chairs, like the
Marquesses in the "Misanthrope"; they whirl round, leap, paint their
faces, gesticulate boldly their ideas; their wit-rockets end with a
song. Young folk, soldiers and artists, they let off their fireworks of
phrases, and gambol round about. "There was a star danced, and
under that was I born."[655] This expression of Beatrice's aptly
describes the kind of poetical, sparkling, unreasoning, charming wit,
more akin to music than to literature, a sort of dream, which is
spoken out aloud, and whilst wide awake, not unlike that described
by Mercutio:
"O, then, I see Queen Mab hath been with you.
She is the fairies' midwife; and she comes
In shape no bigger than an agate-stone
On the fore-finger of an alderman,
Drawn with a team of little atomies
Athwart men's noses as they lie asleep;
Her wagon-spokes made of long spinners' legs,
The cover of the wings of grasshoppers,
The traces of the smallest spider's web,
The collars of the moonshine's watery beams,
Her whip of cricket's bone, the lash of film,
Her wagoner a small gray-coated gnat,
Not half so big as a round little worm
Prick'd from the lazy finger of a maid;
Her chariot is an empty hazel-nut,
Made by the joiner squirrel or old grub,
Time out o' mind the fairies' coachmakers.
And in this state she gallops night by night
Through lovers' brains, and then they dream of love;
O'er courtiers' knees, that dream on court'sies straight,
O'er lawyers' fingers, who straight dream on fees,
O'er ladies' lips, who straight on kisses dream....
Sometime she gallops o'er a courtier's nose,
And then dreams he of smelling out a suit;
And sometime comes she with a tithe-pig's tail
Tickling a person's nose as a' lies asleep,
Then dreams he of another benefice:
Sometime she driveth o'er a soldier's neck,
And then dreams he of cutting foreign throats,
Of breaches, ambuscadoes, Spanish blades,
Of healths five-fathom deep: and then anon
Drums in his ear, at which he starts and wakes,
And being thus frighted swears a prayer or two
And sleeps again. This is that very Mab
That plats the manes of horses in the night,
And bakes the elf-locks in foul sluttish hairs,
Which once untangled much misfortune bodes...
This is she."[656]
CHOICE EXAMPLES OF BOOK ILLUMINATION.
Fac-similes from Illuminated Manuscripts and Illustrated
Books of Early Date.
TITLE-PAGE OF THE HYPNEROTOMACHIA.

The present frontispiece belongs to a French


translation of the work of Poliphilo, the only book
with decorated borders and insertions ever
published by the Venetian Aldi. They printed the
Hypnerotomachia in 1499, and it was reproduced in
a French translation, with the present title-page by
the Parisian printer, Jacques Kerver, in 1546. All the
profuse embellishments of the Aldine edition were
retained, but the title-page here reproduced is from
a design of the famous French sculptor, Jean
Goujon.
Romeo interrupts him, or he would never end. Let the reader
compare with the dialogue of the French theatre this little poem
"Child of an idle brain,
Begot of nothing but vain fantasy,"[657]
introduced without incongruity in the midst of a conversation of the
sixteenth century, and he will understand the difference between the
wit which devotes itself to reasoning, or to record a subject for
laughter, and that imagination which is self-amused with its own act.
Falstaff has the passions of an animal, and the imagination of a man
of wit. There is no character which better exemplifies the fire and
immorality of Shakespeare. Falstaff is a great supporter of
disreputable places, swearer, gamester, idler, wine-bibber, as low as
he well can be. He has a big belly, bloodshot eyes, bloated face,
shaking legs; he spends his life with his elbows among the tavern-
jugs, or asleep on the ground behind the arras; he only wakes to
curse, lie, brag, and steal. He is as big a swindler as Panurge, who
had sixty-three ways of making money, "of which the honestest was
by sly theft." And what is worse, he is an old man, a knight, a
courtier, and well educated. Must he not be odious and repulsive? By
no means; we cannot help liking him. At bottom, like his brother
Panurge, he is "the best fellow in the world." He has no malice in his
composition; no other wish than to laugh and be amused. When
insulted, he bawls out louder than his attackers, and pays them back
with interest in coarse words and insults; but he owes them no
grudge for it. The next minute he is sitting down with them in a low
tavern, drinking their health like a brother and comrade. If he has
vices, he exposes them so frankly that we are obliged to forgive him
them. He seems to say to us, "Well, so I am, what then? I like
drinking: isn't the wine good? I take to my heels when hard hitting
begins; don't blows hurt? I get into debt, and do fools out their
money; isn't it nice to have money in your pocket? I brag; isn't it
natural to want to be well thought of?"—"Dost thou hear, Hal? thou
knowest, in the state of innocency, Adam fell; and what should poor
Jack Falstaff do in the days of villainy? Thou seest I have more flesh
than another man, and therefore more frailty."[658] Falstaff is so
frankly immoral, that he ceases to be so. Conscience ends at a
certain point; nature assumes its place, and man rushes upon what
he desires, without more thought of being just or unjust than an
animal in the neighboring wood. Falstaff, engaged in recruiting, has
sold exemptions to all the rich people, and only enrolled starved and
half-naked wretches. There's but a shirt and a half in all his
company: that does not trouble him. Bah: "they'll find linen enough
on every hedge." The prince, who has seen them, says, "I did never
see such pitiful rascals. Tut, tut," answers Falstaff, "good enough to
toss; food for powder; they'll fill a pit as well as better; tush, man,
mortal men, mortal men."[659] His second excuse is his unfailing
spirit. If ever there was a man who could jabber, it is he. Insults and
oaths, curses, jobations, protests, flow from him as from an open
barrel. He is never at a loss; he devises a shift for every difficulty.
Lies sprout out of him, fructify, increase, beget one another, like
mushrooms on a rich and rotten bed of earth. He lies still more from
his imagination and nature than from interest and necessity. It is
evident from the manner in which he strains his fictions. He says he
has fought alone against two men. The next moment it is four.
Presently we have seven, then eleven, then fourteen. He is stopped
in time, or he would soon be talking of a whole army. When
unmasked, he does not lose his temper, and is the first to laugh at
his boastings. "Gallants, lads, boys, hearts of gold.... What, shall we
be merry? shall we have a play extempore?"[660] He does the
scolding part of King Henry with so much truth that we might take
him for a king, or an actor. This big potbellied fellow, a coward, a
cynic, a brawler, a drunkard, a lewd rascal, a pothouse poet, is one
of Shakespeare's favorites. The reason is, that his morals are those
of pure nature, and Shakespeare's mind is congenial with his own.

SECTION VI.—Shakespeare's Women

Nature is shameless and gross amidst this mass of flesh, heavy with
wine and fatness. It is delicate in the delicate body of women, but as
unreasoning and impassioned in Desdemona as in Falstaff.
Shakespeare's women are charming children, who feel in excess and
love passionately. They have unconstrained manners, little rages,
nice words of friendship, a coquettish rebelliousness, a graceful
volubility, which recall the warbling and the prettiness of birds. The
heroines of the French stage are almost men; these are women, and
in every sense of the word. More imprudent than Desdemona a
woman could not be. She is moved with pity for Cassio, and asks a
favor for him passionately, recklessly, be the thing just or no,
dangerous or no. She knows nothing of man's laws, and does not
think of them. All that she sees is, that Cassio is unhappy:
"Be thou assured, good Cassio... My lord shall never rest;
I'll watch him, tame and talk him out of patience;
His bed shall seem a school, his board a shrift;
I'll intermingle everything he does
With Cassio's suit."[661]
She asks her favor:
"Othello. Not now, sweet Desdemona; some other
time.
Desdemona. But shall't be shortly?
O. The sooner, sweet, for you.
Des. Shall't be to-night at supper?
O. No, not to-night.
Des. To-morrow dinner, then?
O. I shall not dine at home;
I meet the captains at the citadel.
Des. Why, then, to-morrow night; or Tuesday morn;
On Tuesday noon, or night; on Wednesday morn;
I prithee, name the time, but let it not
Exceed three days: in faith, he's penitent."[662]
She is somewhat astonished to see herself refused: she scolds
Othello. He yields: who would not yield seeing a reproach in those
lovely sulking eyes? O, says she, with a pretty pout:
"This is not a boon;
'Tis as I should entreat you wear your gloves,
Or feed on nourishing dishes, or keep you warm,
Or sue to you to do peculiar profit
To your own person."[663]
A moment after, when he prays her to leave him alone for a while,
mark the innocent gayety, the ready observance, the playful child's
tone:
"Shall I deny you? no: farewell, my lord....
Emilia, come: Be as your fancies teach you;
Whate'er you be, I am obedient."[664]
This vivacity, this petulance, does not prevent shrinking modesty and
silent timidity: on the contrary, they spring from a common cause,
extreme sensibility. She who feels much and quickly has more
reserve and more passion than others; she breaks out or is silent;
she says nothing or everything. Such is this Imogen.
"So tender of rebukes that words are strokes,
And strokes death to her."[665]
Such is Virgilia, the sweet wife of Coriolanus; her heart is not a
Roman one; she is terrified at her husband's victories: when
Volumnia describes him stamping on the field of battle, and wiping
his bloody brow with his hand, she grows pale:
"His bloody brow! O Jupiter, no blood!...
Heavens bless my lord from fell Aufidius!"[666]
She wishes to forget all that she knows of these dangers; she dare
not think of them. When asked if Coriolanus does not generally
return wounded, she cries, "O, no, no, no." She avoids this cruel
picture, and yet nurses a secret pang at the bottom of her heart.
She will not leave the house: "I'll not over the threshold till my lord
return."[667] She does not smile, will hardly admit a visitor; she
would blame herself, as for a lack of tenderness, for a moment's
forgetfulness or gayety. When he does return, she can only blush
and weep. This exalted sensibility must needs end in love. All
Shakespeare's women love without measure, and nearly all at first
sight. At the first look Juliet cast on Romeo, she says to the nurse:
"Go, ask his name: if he be married,
My grave is like to be my wedding bed."[668]
It is the revelation of their destiny. As Shakespeare has made them,
they cannot but love, and they must love till death. But this first look
is an ecstasy: and this sudden approach of love is a transport.
Miranda seeing Fernando, fancies that she sees "a thing divine." She
halts motionless, in the amazement of this sudden vision, at the
sound of these heavenly harmonies which rise from the depths of
her heart. She weeps, on seeing him drag the heavy logs; with her
slender white hands she would do the work whilst he reposed. Her
compassion and tenderness carry her away; she is no longer
mistress of her words, she says what she would not, what her father
has forbidden her to disclose, what an instant before she would
never have confessed. The too full heart overflows unwittingly,
happy, and ashamed at the current of joy and new sensations with
which an unknown feeling has flooded her:
"Miranda. I am a fool to weep at what I am glad of....
Fernando. Wherefore weep you?
M. At mine unworthiness that dare not offer
What I desire to give, and much less take
What I shall die to want....
I am your wife, if you will marry me;
If not, I'll die your maid."[669]
This irresistible invasion of love transforms the whole character. The
shrinking and tender Desdemona, suddenly, in full Senate, before
her father, renounces her father; dreams not for an instant of asking
his pardon, or consoling him. She will leave for Cyprus with Othello,
through the enemy's fleet and the tempest. Everything vanishes
before the one and adored image which has taken entire and
absolute possession of her whole heart. So, extreme evils, bloody
resolves, are only the natural sequence of such love. Ophelia
becomes mad, Juliet commits suicide; no one but looks upon such
madness and death as necessary. You will not then discover virtue in
these souls, for by virtue is implied a determinate desire to do good,
and a rational observance of duty. They are only pure through
delicacy or love. They recoil from vice as a gross thing, not as an
immoral thing. What they feel is not respect for the marriage vow,
but adoration of their husband. "O sweetest, fairest lily!" So
Cymbeline speaks of one of these frail and lovely flowers which
cannot be torn from the tree to which they have grown, whose least
impurity would tarnish their whiteness. When Imogen learns that her
husband means to kill her as being faithless, she does not revolt at
the outrage; she has no pride, but only love. "False to his bed!" She
faints at the thought that she is no longer loved. When Cordelia
hears her father, an irritable old man, already almost insane, ask her
how she loves him, she cannot make up her mind to say aloud the
flattering protestations which her sisters have been lavishing. She is
ashamed to display her tenderness before the world, and to buy a
dowry by it. He disinherits her, and drives her away; she holds her
tongue. And when she afterwards finds him abandoned and mad,
she goes on her knees before him, with such a touching emotion,
she weeps over that dear insulted head with so gentle a pity, that
you might fancy it was the tender voice of a desolate but delighted
mother, kissing the pale lips of her child:
"O yon kind gods,
Cure this great breach in his abused nature!
The untuned and jarring senses, O, wind up
Of this child-changed father!...
O my dear father! Restoration hang
Thy medicine on my lips; and let this kiss
Repair those violent harms that my two sisters
Have in thy reverence made!... Was this a face
To be opposed against the warring winds?
... Mine enemy's dog,
Though he had bit me, should have stood that night
Against my fire....
How does my royal lord? How fares your majesty?"[670]
If, in short, Shakespeare comes across a heroic character, worthy of
Corneille, a Roman, such as the mother of Coriolanus, he will explain
by passion what Corneille would have explained by heroism. He will
depict it violent and thirsting for the violent feelings of glory. She will
not be able to refrain herself. She will break out into accents of
triumph when she sees her son crowned; into imprecations of
vengeance when she sees him banished. She will descend to the
vulgarities of pride and anger; she will abandon herself to mad
effusions of joy, to dreams of an ambitious fancy,[671] and will prove
once more that the impassioned imagination of Shakespeare has left
its trace in all the creatures whom it has called forth.
SECTION VII.—Types of Villains

Nothing is easier to such a poet than to create perfect villains.


Throughout he is handling the unruly passions which make their
character, and he never hits upon the moral law which restrains
them; but at the same time, and by the same faculty, he changes
the inanimate masks, which the conventions of the stage mould on
an identical pattern, into living and illusory figures. How shall a
demon be made to look as real as a man? Iago is a soldier of
fortune who has roved the world from Syria to England, who, nursed
in the lowest ranks, having had close acquaintance with the horrors
of the wars of the sixteenth century, had drawn thence the maxims
of a Turk and the philosophy of a butcher; principles he has none
left. "O my reputation, my reputation!" cries the dishonored Cassio.
"As I am an honest man," says Iago, "I thought you had received
some bodily wound; there is more sense in that than in reputation."
[672] As for woman's virtue, he looks upon it like a man who has
kept company with slave-dealers. He estimates Desdemona's love as
he would estimate a mare's: that sort of thing lasts so long—then...
And then he airs an experimental theory with precise details and
nasty expressions like a stud doctor. "It cannot be that Desdemona
should long continue her love to the Moor, nor he his to her.... These
Moors are changeable in their wills;... the food that to him now is as
luscious as locusts, shall be to him shortly as bitter as coloquintida.
She must change for youth: when she is sated with his body, she will
find the error of her choice."[673] Desdemona, on the shore, trying!
to forget her cares, begs him to sing the praises of her sex. For
every portrait he finds the most insulting insinuations. She insists,
and bids him take the case of a deserving woman. "Indeed," he
replies, "she was a wight, if ever such wight were,... to suckle fools
and chronicle small beer."[674] He also says, when Desdemona asks
him what he would write in praise of her: "O gentle lady do not put
me to't, for I am nothing, if not critical."[675] This is the key to his
character. He despises man; to him Desdemona is a little wanton
wench, Cassio an elegant word-shaper, Othello a mad bull, Roderigo
an ass to be basted, thumped, made to go. He diverts himself by
setting these passions at issue; he laughs at it as at a play. When
Othello, swooning, shakes in his convulsions, he rejoices at this
capital result: "Work on, my medicine, work! Thus credulous fools
are caught."[676] You would take him for one of the poisoners of the
time, studying the effect of a new potion on a dying dog. He only
speaks in sarcasms; he has them ready for everyone, even for those
whom he does not know. When he wakes Brabantio to inform him of
the elopement of his daughter, he tells him the matter in coarse
terms, sharpening the sting of the bitter pleasantry, like a
conscientious executioner, rubbing his hands when he hears the
culprit groan under the knife. "Thou art a villain!" cries Brabantio.
"You are—a senator!" answers Iago. But the feature which really
completes him, and makes him take rank with Mephistopheles, is the
atrocious truth and the cogent reasoning by which he likens his
crime to virtue.[677] Cassio, under his advice, goes to see
Desdemona, to obtain her intercession for him; this visit is to be the
ruin of Desdemona and Cassio. Iago, left alone, hums for an instant
quietly, then cries:
"And what's he then that says I play the villain?
When this advice is free I give and honest,
Probal to thinking and indeed the course
To win the Moor again."[678]
To all these features must be added a diabolical energy,[679] an
inexhaustible inventiveness in images, caricatures, obscenity, the
manners of a guard-room, the brutal bearing and tastes of a trooper,
habits of dissimulation, coolness, hatred, and patience, contracted
amid the perils and devices of a military life, and the continuous
miseries of long degradation and frustrated hope; you will
understand how Shakespeare could transform abstract treachery into
a concrete form, and how Iago's atrocious vengeance is only the
natural consequence of his character, life, and training.
SECTION VIII.—Principal Characters

How much more visible is this impassioned and unfettered genius of


Shakespeare in the great characters which sustain the whole weight
of the drama! The startling imagination, the furious velocity of the
manifold and exuberant ideas, passion let loose, rushing upon death
and crime, hallucinations, madness, all the ravages of delirium
bursting through will and reason: such are the forces and ravings
which engender them. Shall I speak of dazzling Cleopatra, who holds
Antony in the whirlwind of her devices and caprices, who fascinates
and kills, who scatters to the winds the lives of men as a handful of
desert dust, the fatal Eastern sorceress who sports with love and
death, impetuous, irresistible, child of air and fire, whose life is but a
tempest, whose thought, ever barbed and broken, is like the
crackling of a lightning flash? Of Othello, who, beset by the graphic
picture of physical adultery, cries at every word of Iago like a man
on the rack; who, his nerves hardened by twenty years of war and
shipwreck, grows mad and swoons for grief, and whose soul,
poisoned by jealousy, is distracted and disorganized in convulsions
and in stupor? Or of old King Lear, violent and weak, whose half-
unseated reason is gradually toppled over under the shocks of
incredible treacheries, who presents the frightful spectacle of
madness, first increasing, then complete, of curses, bowlings,
superhuman sorrows, into which the transport of the first access of
fury carries him, and then of peaceful incoherence, chattering
imbecility, into which the shattered man subsides; a marvellous
creation, the supreme effort of pure imagination, a disease of
reason, which reason could never have conceived?[680] Amid so
many portraitures let us choose two or three to indicate the depth
and nature of them all. The critic is lost in Shakespeare, as in an
immense town; he will describe a couple of monuments, and entreat
the reader to imagine the city.
Plutarch's Coriolanus is an austere, coldly haughty patrician, a
general of the army. In Shakespeare's hands he becomes a coarse
soldier, a man of the people as to his language and manners, an
athlete of war, with a voice like a trumpet; whose eyes by
contradiction are filled with a rush of blood and anger, proud and
terrible in mood, a lion's soul in the body of a bull. The philosopher
Plutarch told of him a lofty philosophic action, saying that he had
been at pains to save his landlord in the sack of Corioli.
Shakespeare's Coriolanus has indeed the same disposition, for he is
really a good fellow; but when Lartius asks him the name of this
poor Volscian, in order to secure his liberty, he yawns out:
"By Jupiter! forgot.
I am weary; yea, my memory is tired.
Have we no wine here?"[681]
He is hot, he has been fighting, he must drink; he leaves his Volscian
in chains, and thinks no more of him. He fights like a porter, with
shouts and insults, and the cries from that deep chest are heard
above the din of the battle like the sounds from a brazen trumpet.
He has scaled the walls of Corioli, he has butchered till he is gorged
with slaughter. Instantly he turns to the army of Cominius, and
arrives red with blood, "as he were flay'd. Come I too late?"
Cominius begins to compliment him. "Come I too late?" he repeats.
The battle is not yet finished: he embraces Cominius:
"O! let me clip ye
In arms as sound as when I woo'd, in heart
As merry as when our nuptial day was done."[682]
For the battle is a real holiday to him. Such senses, such a strong
frame, need the outcry, the din of battle, the excitement of death
and wounds. This haughty and indomitable heart needs the joy of
victory and destruction. Mark the display of his patrician arrogance
and his soldier's bearing, when he is offered the tenth of the spoils:
"I thank you, general;
But cannot make my heart consent to take
A bribe to pay my sword."[683]
Welcome to Our Bookstore - The Ultimate Destination for Book Lovers
Are you passionate about books and eager to explore new worlds of
knowledge? At our website, we offer a vast collection of books that
cater to every interest and age group. From classic literature to
specialized publications, self-help books, and children’s stories, we
have it all! Each book is a gateway to new adventures, helping you
expand your knowledge and nourish your soul
Experience Convenient and Enjoyable Book Shopping Our website is more
than just an online bookstore—it’s a bridge connecting readers to the
timeless values of culture and wisdom. With a sleek and user-friendly
interface and a smart search system, you can find your favorite books
quickly and easily. Enjoy special promotions, fast home delivery, and
a seamless shopping experience that saves you time and enhances your
love for reading.
Let us accompany you on the journey of exploring knowledge and
personal growth!

ebookgate.com

You might also like