Elementary Vector and Tensor Analysis - Brannon
Elementary Vector and Tensor Analysis - Brannon
Elementary Vector and Tensor Analysis - Brannon
R. M. Brannon
University of New Mexico, Albuquerque
Copyright is reserved.
FT
University
Abstract
RA
Elementary vector and tensor analysis concepts are reviewed using a notation
that proves useful for higher-order tensor analysis of anisotropic media.
Acknowledgments
RA
FT
ii
Contents
ii
xiii
1
3
7
7
8
9
9
10
10
10
10
11
11
13
14
17
17
18
19
19
20
22
23
24
24
24
25
25
26
27
29
29
30
31
31
32
33
34
34
35
35
36
RA
FT
Acknowledgments ...............................................................................................
Preface.....................................................................................................................
Introduction ............................................................................................................
Terminology from functional analysis .............................................................
Matrix Analysis.......................................................................................................
Definition of a matrix...........................................................................................
The matrix product...............................................................................................
The transpose of a matrix.....................................................................................
The inner product of two column matrices ..........................................................
The outer product of two column matrices. .........................................................
The trace of a square matrix.................................................................................
The Kronecker delta.............................................................................................
The identity matrix...............................................................................................
The 3D permutation symbol ................................................................................
The - (E-delta) identity.....................................................................................
The - (E-delta) identity with multiple summed indices ...................................
Determinant of a square matrix............................................................................
Principal sub-matrices and principal minors........................................................
Matrix invariants..................................................................................................
Positive definite ...................................................................................................
The cofactor-determinate connection ..................................................................
Inverse..................................................................................................................
Eigenvalues and eigenvectors ..............................................................................
Similarity transformations .............................................................................
Finding eigenvectors by using the adjugate.........................................................
Vector/tensor notation........................................................................................
Ordinary engineering vectors ...........................................................................
Engineering laboratory base vectors ................................................................
Other choices for the base vectors .......................................................................
Basis expansion of a vector..................................................................................
Summation convention details........................................................................
Dont forget to remember what repeated indices really mean ......................
Indicial notation in derivatives ......................................................................
BEWARE: avoid implicit sums as independent variables..............................
Reading index STRUCTURE, not index SYMBOLS .........................................
Aesthetic (courteous) indexing ............................................................................
Suspending the summation convention ...............................................................
Combining indicial equations ..............................................................................
The index-changing property of the Kronecker delta ..........................................
Summing the Kronecker delta itself ....................................................................
The under-tilde notation ...................................................................................
Simple vector operations and properties ......................................................
Dot product between two vectors ........................................................................
Dot product between orthonormal base vectors...................................................
iii
36
37
37
38
38
39
39
41
41
42
43
44
44
46
47
47
48
49
49
50
RA
FT
51
53
53
53
54
55
57
58
61
62
63
64
64
64
65
66
66
69
69
69
70
71
71
71
73
iv
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
73
76
76
77
78
79
80
80
82
82
83
83
84
84
85
85
86
87
89
89
89
91
93
94
95
96
97
97
98
98
99
99
101
101
102
102
104
104
108
108
111
111
112
114
116
116
RA
FT
117
118
118
119
120
121
121
122
122
122
122
123
125
126
128
128
129
130
130
131
132
132
132
133
135
135
136
138
139
140
140
142
143
144
146
146
147
147
149
150
151
151
155
155
157
158
RA
FT
Primitive invariants..............................................................................................
Trace invariants....................................................................................................
Characteristic invariants.......................................................................................
Direct notation definitions of the characteristic invariants...........................
The cofactor in the triple scalar product .......................................................
Invariants of a sum of two tensors .......................................................................
CASE: invariants of the sum of a tensor plus a dyad ....................................
The Cayley-Hamilton theorem: ...........................................................................
CASE: Expressing the inverse in terms of powers and invariants.................
Inverse of the sum of a tensor plus a dyad...........................................................
Eigenvalue problems............................................................................................
Algebraic and geometric multiplicity of eigenvalues ..........................................
Diagonalizable tensors .........................................................................................
Eigenprojectors ....................................................................................................
Geometrical entities ............................................................................................
Equation of a plane ..............................................................................................
Equation of a line .................................................................................................
Equation for a sphere ...........................................................................................
Equation for an ellipsoid......................................................................................
Example..........................................................................................................
Equation for a cylinder with an ellipse-cross-section ..........................................
Equation for a right circular cylinder...................................................................
Equation of a general quadric (including hyperboloid) .......................................
Generalization of the quadratic formula and completing the square................
Polar decomposition............................................................................................
The Q-R decomposition.......................................................................................
The polar decomposition theorem: ......................................................................
The polar decomposition is a nonlinear projection operation .............................
The *FAST* way to do a polar decomposition in two dimensions.....................
Material symmetry................................................................................................
Isotropic second-order tensors in 3D space .........................................................
Isotropic second-order tensors in 2D space .........................................................
Isotropic fourth-order tensors ..............................................................................
Transverse isotropy..............................................................................................
Abstract vectors/tensor algebra......................................................................
Definition of an abstract vector............................................................................
Inner product spaces ............................................................................................
Continuous functions are vectors! .......................................................................
Tensors are vectors! .............................................................................................
Vector subspaces..................................................................................................
Subspaces and the projection theorem.................................................................
Abstract contraction and swap (exchange) operators ..........................................
The contraction tensor ...................................................................................
The swap tensor .............................................................................................
Vector/tensor calculus........................................................................................
ASIDE #1: total and partial derivative notation...................................
vi
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
160
160
163
164
166
167
169
171
175
RA
FT
vii
FT
RA
D
viii
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
Figures
FT
RA
Figure 5.1.
Figure 5.2.
Figure 6.1.
Figure 6.2.
Figure 6.3.
Figure 6.4.
Figure 6.5.
Figure 6.6.
Figure 6.7.
Figure 17.1.
ix
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
FT
RA
D
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
RA
FT
Tables
xi
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
FT
RA
D
xii
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
Preface
Math and science journals often have extremely restrictive page limits,
making it virtually impossible to present a coherent development of complicated concepts by working upward from basic concepts. Furthermore, scholarly journals are intended for the presentation of new results, so detailed
explanations of known results are generally frowned upon (even if those
results are not well-known or well-understood). Consequently, only those
readers who are already well-versed in a subject have any hope of effectively
reading the literature to further expand their knowledge. While this situation
is good for experienced researchers and specialists in a particular field of
study, it can be a frustrating handicap for less experienced people or people
whose expertise lies elsewhere. This report serves these individuals by presenting several known theorems or mathematical techniques that are useful
for the analysis material behavior. Most of these theorems are scattered willynilly throughout the literature. Several rarely appear in elementary textbooks. Most of the results in this report can be found in advanced textbooks on
functional analysis, but these books tend to be overly generalized, so the application to specific problems is unclear. Advanced mathematics books also tend
to use notation that might be unfamiliar to the typical research engineer. This
report presents derivations of theorems only where they help clarify concepts.
The range of applicability of theorems is also omitted in certain situations. For
example, describing the applicability range of a Taylor series expansion
requires the use of complex variables, which is beyond the scope of this document. Likewise, unless otherwise stated, I will always implicitly presume that
functions are well-behaved enough to permit whatever operations I perform.
For example, the act of writing df dx will implicitly tell the reader that I am
assuming that f can be written as a function of x and (furthermore) this function is differentiable. In the sense that much of the usual (but distracting)
mathematical provisos are missing, I consider this document to be a work of
engineering despite the fact that it is concerned principally with mathematics.
While I hope this report will be useful to a broader audience of readers, my
personal motivation is to establish a single bibliographic reference to which I
can point from my more stilted and terse journal publications.
Rebecca Brannon,
rmbrann@sandia.gov
Sandia National Laboratories
September 10, 2002 11:51 am.
xiii
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
xiv
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
1. Introduction
This report reviews tensor algebra (and a bit of tensor calculus) using a
notation that proves very useful when extending these basic ideas to higher
dimensions. Tensor notation unfortunately remains non-standardized, so its
important to at least scan this report to become familiar with our definitions
and overall approach to the field if you wish to move along to our other (more
sophisticated and contemporary) applications in materials modeling.
Many of the definitions and theorems provided in this report are not rigorous instead they are presented in a more physical engineering manner.
More careful expositions on these topics can be found in elementary textbooks
on matrix, vector, and tensor analysis. One may need to look to a functional
analysis text to find an equally developed discussion of projections. The reader
is presumed to have been exposed to vector analysis and matrix analysis at
the rudimentary level that is ordinarily covered in undergraduate calculus
courses.
1
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
We present the information in this report in order to have a single reference in which all the concepts are presented together using a notation that is
consistent with that used in more advanced follow-up work that we intend to
published separately. Some of our other work explains that many theorems in
higher-dimensional realms have perfect analogs with the ordinary concepts
from 3D. For example, this elementary report discusses how to obliquely
project a vector onto plane, and we demonstrate in later (more advanced) work
that the act of solving viscoplasticity models by a return mapping algorithm is
perfectly analogous to vector projection.
Throughout this report, we use the term ordinary to refer to the three
dimensional physical space in which everyday engineering problems occur.
The term abstract will be used later when extending ordinary concepts to
higher dimensional spaces, which is the principal goal of generalized tensor
analysis. Except where otherwise stated, the basis { e 1, e 2, e 3 } used for vectors
orthonormal
and tensors in this report will be assumed regular (i.e.,
and righthanded). Thus, all indicial formulas in this report use what Malvern [12] calls
rectangular Cartesian components. Readers interested in irregular bases can
find a discussion of curvilinear coordinates at http://www.me.unm.edu/
~rmbrann/gobag.html (however, that report presumes that the reader is
already familiar with the notation and basic identities that are covered in this
report).
2
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
* At this point, the reader is not expected to already know what is meant by the term tensor, much
less the order of a tensor or the meaning of inner product. For now, consider this section to
apply to scalars and vectors. Just understand that the concepts reviewed in this section will also
apply in more general tensor settings, once learned.
3
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
4
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
5
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
n 12 + n 22 + n 32 = 1 .
If a set is closed under vector addition and scalar multiplication (i.e., if
every linear combination of set members gives a result that is also in the
set), then the set is called a linear manifold, or a linear space. Otherwise,
the set is curvilinear. The set of all unit vectors is a curvilinear space
because a linear combination of unit vectors does not result in a unit
vector. Linear manifolds are like planes that pass through the origin,
though they might be hyperplanes, which is just a fancy word for a plane
of more than just two dimensions. Linear spaces can also be onedimensional. Any straight line that passes through the origin is a linear
manifold.
Zero must always be a member of a linear manifold, and this fact is often
a great place to start when considering whether or not a set is a linear
space. For example, we could assert that the set of unit vectors is not a
linear space by simply noting that the zero vector is not a unit vector.
A plane that does not pass through the origin must not be a linear space.
We know this simply because such a plane does not contain the zero
vector. This kind of plane is called an affine space. An affine space is a
set that would become a linear space if the origin were to be moved to any
single point in the set. For example, the point ( 0, b ) lies on the straight
line defined by the equation, y = mx + b . If we move the origin from
O = ( 0, 0 ) to a new location O * = ( 0, b ) , and introduce a change of
variables x * = x 0 and y * = y b , then the equation for this same line
described with respect to this new origin would become y * = mx * , which
does describe a linear space. Stated differently, a set S is affine if every
member x in that set is expressible in the form of a constant vector d plus
a vector x * that does belong to a linear space. Thus, learning about the
properties of linear spaces is sufficient to learn most of what you need to
know about affine spaces.
Given an n-dimensional linear space, a subset of members of that space is
basis if every member of the space can be expressed as a linear
combination of members of the subset. A basis always contains exactly as
many members as the dimension of the space.
A binary operation is simply a function or transformation that has two
arguments. For example, f ( x, y ) = x 2 y is a binary operation.
6
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Definition of a matrix
A matrix is an ordered array of numbers that are typically arranged in the
form of a table having N rows and M columns. If one of the dimensions ( N
or M ) happens to equal 1, then the term vector is often used, although we
will often use the term array in order to avoid confusion with vectors in the
physical sense. A matrix is called square if M= N . We will usually typeset
matrices in plain text with brackets such as [ A ] .
For matrices of dimension N 1 , we may also use braces, as in { v } ;
namely, if N =3 , then
v1
{ v } = v2
(3.1)
v3
For matrices of dimension 1 M , we use angled brackets <v> ; Thus, if N =3 ,
then
<v> = [ v 1, v 2, v 3 ]
cca B
ranno
(3.2)
7
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
In this report, vectors will be typeset in bold with one single under-tilde
(for example, v ) and the associated three components of the vector with
implicitly understood basis will be denoted { v } or <v > ,
respect to some
(3.3)
MN
= [ A]
MR
[ B]
(3.4)
RN
Note that the dimension R must be common to both matrices on the righthand side, and this common dimension must reside at the inside or abutting position (the trailing dimension of [ A ] must equal the leading dimension of [ B ] )
The matrix product operation is defined
R
C ij =
A ik B kj ,
k=1
(3.5)
(3.6)
ui =
F ik v k ,
(3.7)
k=1
8
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
MN
is a new matrix [ B ]
NM
(note the
B ij = A ji
where i takes values from 1 to N ,
and j takes values from 1 to M .
(3.8)
ij
A ijT = A ji
where i takes values from 1 to N ,
and j takes values from 1 to M .
(3.9)
The dimensions of [ A ] and [ A ] T are reverses of each other. Thus, for example, if { v } is an N 1 matrix, then { v } T is a 1 N matrix. In other words,
{ v } T = <v>
and
<v> T = { v }
(3.10)
The transpose of a product is the reverse product of the transposes. For example,
( [ A ] [ B ] ) T = [ B ] T [ A ] T , and
( <v> [ A ] ) T = [ A ] T <v> T = [ A ] T { v }
(3.11)
(3.12)
vk wk
cca B
ranno
(3.13)
k=1
product gives the same result as the vector dot product v w , defined later.
9
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(3.14)
For this case, the value of the adjacent dimension R in Eq. (3.5) is just 1, so
the summation ranges from 1 to 1 (which means that it is just a solitary term).
The result of the outer product is an M N matrix, whose ij component is
given by a i b j . If { a } and { b } contain components of two vectors a and b
then the outer product gives the matrix corresponding to the dyadic product
a b (also often denoted a b ), to be discussed in gory detail later.
tr [ A ] =
A kk
(3.15)
k=1
(3.16)
(cyclic property)
(3.17)
(3.18)
(3.19)
The ij component of the identity is given by ij . Note that, for any array { v }
[ I ]{v} = {v}
(3.20)
10
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
ijk
otherwise
0
+1
(3.21)
Note that the indices may be permuted cyclically without changing the value
of the result. Furthermore, inverting any two indices will change the sign of
the value. Thus, the permutation symbol has the following properties:
ijk = jki = kij = jik = ikj = kji
(3.22)
The term 3D is used to indicate that there are three subscripts on ijk each
of which take on values from 1 to 3.*
ijn kln
= ik jl il jk .
(3.23)
n=1
cca B
ranno
(3.24)
* Though not needed for our purposes, the 2D permutation symbol ij is defined to equal zero if i= j ,
+1 if ij=12 , and 1 if ij = 21 . The 4D permutation symbol ijkl is defined to equal zero if any of
the four indices are equal; it is +1 if ijkl is an even permutation of 1234 and 1 if ijkl is an odd
permutation. A permutation is simply a rearrangement. The permutation ijkl is even if rearranging
it back to 1234 can be accomplished by an even number of moves that exchange two elements at a
time. A cyclic permutation of an n-D permutation symbol will change sign if n is even, but remain
unchanged if n is odd. Thus, for our 3D permutation symbol, cyclic permutations dont change
sign, whereas cyclic permutations of the 4D permutation symbol will change sign.
11
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
To make an expression fit the index structure of Eq. (3.23), most people
laboriously apply the cyclic property to each alternating symbol until the
summed index is located at the trailing side on both of them. Keeping track of
whether or not these manipulations will require changing the final sign of the
right hand side of the - identity is one of the most common and avoidable
careless mistakes made when people use this identity. Even once the summation index has been properly positioned at the trailing end of each alternating
symbol, most people then apply a slow (and again error-prone process of figuring out where the free indices go). Typically people apply a left-right/outsideinside rule. By this, we mean that the free indices on the left sides of ijn and
kln are the indices that go on the first , then the right free indices go on the
second , then the outer free indices go on the third , and (finally) the inner
free indices go on the last . The good news is... you dont have to do it this
way! By thinking about the - identity in a completely different way, you can
avoid both the initial rearrangement of the indices on the alternating symbols
and the slow left-right-out-in placement of the indices. Lets suppose you want
to apply the - identity to the expression imk pin . First write a skeleton of
the identity as follows
imk pin = ?? ?? ?? ??
(3.25)
Our goal is to find a rapid and error-minimizing way to fill in the question
marks with the correct index symbols. Once you have written the skeleton,
look at the left-hand side to identify which index is summed. In this case, it is
the index i . Next say out loud the four free indices in an order defined by
cyclically moving forward from the summed index on each alternating symbol. Each alternating symbol has two free indices. To call out their names by
moving cyclically forward, you simply say the name of the two indices to the
right of the summed index, wrapping back around to the beginning if necessary. For example, the two indices cyclically forward from p in the sequence
pqr are qr; the two indices cyclically forward from q are rp; the two
indices forward from r are pq. For the first alternating symbol in our skeleton of Eq. (3.25), the two indices cyclically forward from the summed index i
are mk whereas the two indices cyclically forward from i in the second alternating symbol are np. You can identify these pairs quickly without ever hav-
12
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
ing to rearrange anything, and you can (in your head) group the pairs together
to obtain a sequence of four free indices mknp. The final step is to write
these four indices onto the skeleton. If the indices are ordered 1234, then you
should write the first two indices (first and second) on the skeleton like this
1? 2? 1? 2?
(3.26)
You write the last pair (third and fourth) in order (34) on the first term and in
reverse order (43) on the last term:
13 24 14 23
(3.27)
Thus, for example, to place the free indices mknp onto the Kronecker deltas,
you would first take care of the mk by writing
m? k? m? k?
(3.28)
Then you just finish off with the last two np free indices by writing them
first in that order on the first term and in reverse order on the second term to
obtain the final result:
imk pin = mn kp mp kn .
(3.29)
This may seem a bit strange at first (especially if you are already stuck in the
left-right-out-in mind set), but this method is far quicker and less error-prone.
Give it a try until you become comfortable with it, and you probably wont
dream of going back to your old way.
ijn kln
= ik jl il jk .
(3.30)
n=1
What happens if we now consider the case of two alternating symbols multiplied side-by-side with two indices being summed? This question is equivalent
to throwing a summation around the above equation in such a manner that
we add up only those terms for which j=l . Then
3
ijn kjn
j = 1n = 1
( ik jj ij jk )
j=1
= ik ( 11 + 22 + 33 )
= 3 ik ik
= 2 ik
cca B
ranno
( i1 1k + i2 2k + i3 3k )
(3.31)
T
F
A
R
D
cca
Rebe
Brann
on
Using similar logic, the - identity with all indices summed is equivalent
to setting i=k in the above equation, summing over each instance so that the
result is six. To summarize using the summation conventions,
ijn kjn = 2 ik
(3.32)
ijk ijk = 6
(3.33)
(3.34)
A 11 A 12
A 21 A 22
A 11 A 22 A 12 A 21
(3.35)
( A 11 A 22 A 33 + A 12 A 23 A 31 + A 13 A 21 A 32 )
( A 13 A 22 A 31 + A 11 A 23 A 32 + A 12 A 21 A 33 )
(3.36)
Note that we have arranged this formula such that the first indices in each
factor are 123. For the positive terms, the second indices are all the positive
permutations of 123. Namely: 123, 231, and 312. For the negative terms, the
second indices are all the negative permutations of 123. Namely: 321, 132, and
213. This relationship may be written compactly by using the permutation
symbol ijk from Eq. (3.21). Namely, if [ A ] is a 3 3 matrix, then
14
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
det [ A ] =
Rebe
3
(3.37)
i = 1 j = 1k = 1
This definition can be extended to square matrices of arbitrarily large dimension by using the n-dimensional permutation symbol (see footnote on page 11).
Alternatively, for square matrices of arbitrarily large dimension, the determinant can be defined recursively as
N
det [ A ] N N =
A ij A ijC
i)
(3.38)
j=1
where i is a free index taking any convenient value from 1 to N (any choice
for i will give the same result). The quantity A ijC is called the cofactor of
A ij , and it is defined by
A ijC = ( 1 ) i + j det [ M ij ] ( N 1 ) ( N 1 )
(3.39)
Here [ M ij ] is the submatrix obtained by striking out the i th row and j th column of [ A ] . The determinant of [ M ij ] is called the minor associated with
A ij . By virtue of the ( 1 ) i + j , the cofactor component A ijC is often called the
signed minor. The formula in Eq. (3.38) is almost never used in numerical
calculations because it requires too many multiplications,* but it frequently
shows up in theoretical analyses.
The index i in Eq. (3.39) may be chosen for convenience (usually a row
with several zeros is chosen to minimize the number of sub-determinants that
must be computed). The above definition is recursive because det [ A ] N N is
defined in terms of smaller ( N 1 ) ( N 1 ) determinants, which may in turn
be expressed in terms of ( N 2 ) ( N 2 ) determinants, and so on until the
determinant is expressed in terms of only 1 1 determinants, for which the
determinant is defined in Eq. (3.34). As an example, consider using Eq. (3.38)
to compute the determinant of a 3 3 matrix. Choosing i=1 , Eq. (3.38) gives
A 11 A 12 A 13
det A 21 A 22 A 23 = A 11 det
A 31 A 32 A 33
A 22 A 23
A 32 A 33
A 12 det
A 21 A 23
A 31 A 33
+ A 13 det
A 21 A 22
A 31 A 32
cca B
ranno
(3.40)
15
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A 11 A 12 A 13
det A 21 A 22 A 23 = A 21 det
A 31 A 32 A 33
A 11 A 13
A 32 A 33
+ A 22 det
A 11 A 13
A 31 A 33
A 23 det
A 11 A 12
A 31 A 32
(3.41)
After using Eq. (3.35) to compute the 2 2 submatrices, both of the above
expressions give the same final result as Eq. (3.36).
Some key properties of the determinant are listed below:
det ( [ A ] T ) = det [ A ]
(3.42)
(3.43)
det ( [ A ] N N ) = N det [ A ]
(3.44)
1
det ( [ A ] 1 ) = ----------------det [ A ]
(3.45)
pqr det [ A ] =
(3.48)
i = 1 j = 1k = 1
or, using the summation convention in which repeated indices are understood
to be summed (and, for clarity, now shown in red),
pqr det [ A ] = ijk A pi A qj A rk
(3.49)
(3.50)
Here, there are implied summations over the indices i,j,k,p,q, and r. If it were
expanded out, the above expression would contain 729 terms, so it is obviously
not used to actually compute the determinant. However, it is not at all uncommon for expressions like this to show up in analytical analysis, and it is therefore essential for the analyst to recognize that the right-hand-side simplifies
so compactly.
16
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(3.51)
A 31 A 33
A 12 A 13
Matrix invariants
The k th characteristic invariant, denoted I k , of a matrix [ A ] is the sum
of all possible k k principal minors. For a 3 3 matrix, these three invariants are
I 1 = A 11 + A 22 + A 33
I 2 = det
A 11 A 12
A 21 A 22
+ det
(3.52)
A 11 A 13
A 31 A 33
+ det
A 22 A 23
(3.53)
A 32 A 33
A 11 A 12 A 13
I 3 = det A 21 A 22 A 23
cca B
ranno
(3.54)
A 31 A 32 A 33
Warning: if the matrix is non-symmetric, the characteristic invariants are not
a complete set of independent invariants. If all three characteristic invariants
of a symmetric matrix are zero, then the matrix itself is zero. However, as discussed later, it is possible for all three characteristic invariants of a non-symmetric matrix to be zero without the matrix itself being zero.
17
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Positive definite
A square matrix [ B ] N N is positive definite if and only if
{ v }T [ B ]{ v } > 0
for all { v }
In indicial notation, this requirement is
N
(3.55)
vi Bij v j > 0
(3.56)
i=1j=1
(3.57)
B 12 + B 21
2 ------------------------- v 2 v 1 .
Note that the middle two terms can be combined and written as
B 11 + B 11
- . The third term can also be
Similarly, we can write the first term as 2 -----------------------
2
so written. Thus, the requirement for positive definiteness depends only on
the symmetric part of the matrix [ B ] . The non-symmetric part has no influence on whether or not a matrix is positive definite. Consequently, we may
replace Eq. (3.55) by the equivalent, but more carefully crafted statement:
It can be shown that, a matrix is positive definite if and only if the characteristic invariants of the symmetric part of the matrix are all positive.*
Fortunately, there is an even simpler test for positive definiteness: you only
have to verify that any nested set of principal minors are all positive! This calculation is easier than finding the invariants themselves because it requires
evaluation of only one principal minor determinant of each size (you dont
have to evaluate all of them).
* It is possible to construct a matrix that has all positive invariants, but whose symmetric part does
not have all positive invariants.
18
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(3.58)
By virtue of Eq. (3.42), we note that the transpose of the cofactor matrix is
identically equal to the cofactor matrix associated with [ A ] T . In other words,
the cofactor and transpose operations commute:
( [ A ]C )T = ( [ A ]T )C
(3.59)
if
0
A ik A Cjk =
det [ A ] if
k=1
i j
i= j
(3.60)
A ik A Cjk = det [ A ] ij
(3.61)
k=1
(3.62)
It turns out that the location of the transpose and cofactor operations is inconsequential the result will be the same in all cases. Namely, the following
results also hold true:
[ A ] [ A ] CT = [ A ] C [ A ] T = [ A ] T [ A ] C = [ A ] CT [ A ] = ( det [ A ] ) [ I ]
(3.63)
Inverse
The inverse of a matrix [ A ] is the matrix denoted [ A ] 1 for which
[ A ] [ A ] 1 = [ A ] 1 [ A ] = [ I ]
(3.64)
If the inverse exists, then it is unique. If the inverse does not exist, then the
matrix [ A ] is said to be non-invertible or singular. A necessary and sufficient condition for the inverse to exist is that the determinant must be nonzero:
det [ A ]
cca B
ranno
(3.65)
19
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Comparing Eqs. (3.63) and (3.64), we note that the inverse may be readily
computed from the cofactor by
[ A ] CT
[ A ] 1 = ----------------det [ A ]
(3.66)
While this definition does uniquely define the inverse, it must never be used
as a definition of the cofactor matrix. The cofactor matrix is well-defined and
generally nonzero even if the matrix [ A ] is singular.
I 2 = det
A 11 A 12
A 21 A 22
(3.67)
A 11 A 12
A 21 A 22
+ det
A 11 A 13
A 31 A 33
+ det
A 22 A 23
A 32 A 33
, and
A 11 A 12 A 13
(3.68)
I 3 = det A 21 A 22 A 23
A 31 A 32 A 33
20
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(no sum on i ).
(3.69)
The solution for { p } i will have an undetermined magnitude and, for symmetric matrices, it is conventional to set the magnitude to one. For non-symmetric
matrices, however, the normalization convention is different, as explained
below.
If an eigenvalue i has algebraic multiplicity m (i.e., if the characteristic
equation gives a root i repeated m times), then there can be no more than a
total of m independent eigenvectors associated with that eigenvalue there
might be fewer (though there is always at least one). If the matrix [ A ] is symmetric, then it is well known [22] that it is always possible to find m independent eigenvectors. The directions of the eigenvectors when the multiplicity m
is greater than one are arbitrary. However, the one thing that is unique is the
span of these vectors (see page 5), and it is conventional to set the eigenvectors to any orthonormal set of vectors lying in the span. For non-symmetric
matrices, it might happen that an eigenvalue of multiplicity m corresponds to
a total of < m linearly independent eigenvectors, where is called the geometric multiplicity. For example, the matrix
53
05
(3.70)
Has an eigenvalue = 5 with algebraic multiplicity of two. To find the associate eigenvector(s), we must solve
5 3 p1 = 5 p1
p2
0 5 p2
cca B
ranno
(3.71)
(3.72)
5 p2 = 5 p2
(3.73)
The second equation gives us no information, and the first equation gives the
constraint that p 2 = 0 . Therefore, we have only one eigenvector (geometric
multiplicity equals one) given by { 1, 0 } even though the eigenvalue had algebraic multiplicity of two. When the geometric multiplicity of an eigenvector is
less than the algebraic multiplicity, then there does still exist a subspace that
is uniquely associated with the multiple eigenvalue. However, characterizing
this subspace requires solving a generalized eigenproblem to construct additional vectors that will combine with the one or more ordinary eigenvectors to
21
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
form a set of vectors that span the space. The process for doing this is onerous,
and we have not yet personally happened upon any engineering application
for which finding these generalized eigenvectors provides any useful information, so we will not cover the details. Instructions for the process can be found
in [23,21, 24]. If the generalized eigenvectors are truly sought, then they can
be found via the JordanDecomposition command in Mathematica [25] (see
discussion below to interpret the result).
Similarity transformations. Suppose that we have a set of eigenvalues
{ 1, 2, , N } for a matrix [ A ] , possibly with some of these eigenvalues having algebraic multiplicities greater than one. Let [ L ] denote the matrix whose
columns contain the corresponding eigenvectors (augmented, where necessary, to include generalized eigenvectors for the cases where the geometric
multiplicity is less than the algebraic multiplicity; the ordinary eigenvectors
corresponding to a given eigenvalue should always, by convention, be entered
into columns of [ L ] before the generalized eigenvectors). Then it can be shown
that the original matrix [ A ] satisfies the similarity transformation
[ A ] = [ L ] [ ] [ L ] 1
(3.74)
(3.75)
This result can be obtained in Mathematica [25] via the command JordanDecomposition[{{5,3},{0,5}}]. From this result, we note that the presence of
the 1 in the 12 position of the [ ] matrix implies that the second column of
[ L ] must contain a generalized eigenvector.
The matrix L will be orthogonal (i.e., [ L ] 1 = [ L ] T ) if and only if the original matrix [ A ] is symmetric. For symmetric matrices, there will never be any
generalized eigenvectors (i.e., the algebraic and geometric eigenvalue multiplicities will always be equal), and the [ ] matrix will therefore always be
fully diagonal (no 1 on any off-diagonal).
22
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(3.78)
100
[ B ] CT = 0 0 0
000
1
0
0
cca B
ranno
, and therefore
(3.79)
23
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
24
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
v = v1 e1 + v2 e2 + v3 e3 =
vi e i
(4.1)
i=1
where v i are called the components of the vector with respect to the basis.
The zero vector 0 is defined to be the vector whose components are all zero.
(4.2)
v3
and the expansion of Eq. (4.1) is analogous to writing
v1
1
0
0
v2 = v1 0 + v2 1 + v3 0
0
0
1
v3
(4.3)
cca B
ranno
(4.4)
25
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
{ v } T = <v> ,
(4.5)
(4.6)
The summation convention is defined such that any index (in this case i) that
is repeated exactly twice in a term is understood to be summed from 1 to 3.
The summation ranges up to 3 because ordinary engineering vectors are
always referenced to 3D physical space.
Later on, we will define quantities (like matrix components A ij ) that have
more that one index. Then, for example, the expression A ij v j , for which the
index j is repeated, would mean the same thing as
3
A ij v j
(4.7)
j=1
In this expression, note that the index i occurs exactly once and is not
repeated. Therefore the above expression is actually three expressions, corresponding to the index i taking the values from 1 to 3. For rectangular Cartesian components, the summation convention has two fundamental rules
(extra rules that apply for irregular bases can be found in Ref. [4]):
1. An index that occurs exactly twice in a term is called a dummy index, and
it is understood to be summed from 1 to 3, with the implied summation
symbol applied only to the term in which the repeated dummy index
appears.
2. An index that occurs exactly once in a term is called a free index, and it
must also appear exactly once in every other term.
The following expressions violate the summation convention:
ai + b j
(violates rule 2)
a i A ij b i
(violates rule 1)
(4.8)
26
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(4.9a)
(4.9b)
(4.9c)
(4.9d)
k=1
( b k A ik ) + a i +
( w jk U pkqi T jp vq )
(4.10)
j = 1k = 1 p = 1q = 1
( bk Aik ) + ai + ( w jk U pkqi T jp vq )
k=1
cca B
ranno
(4.11)
q = 1k = 1 p = 1 j = 1
Moving the summation symbols from the jkpq order to this qkpj order has no
impact on the result.
Dont forget to remember what repeated indices really mean. Beginning students sometimes forget to recall that the summation rules are really
just a notational convenience. Sometimes its wise to go back to conventional
notation to simplify an indicial expression. Recall, for example, the definition
of the Kronecker delta:
27
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
1 if i= j
ij =
0 if i j
(4.12)
(4.13)
There are no free indices, so the result must be a scalar. New students almost
always get burned by using Eq. (4.12) to say that kk must be equal to 1. However, this conclusion is wrong. The index k is repeated, so it must be summed.
In other words, Eq. (4.13) really says
3
s =
kk
= 11 + 22 + 33 = 1 + 1 + 1 = 3
(4.14)
k=1
Recalling the true meaning of the indicial notation is also essential for simplifying other expressions where ij appears. Consider, for example,
A km mj
(4.15)
A km mj
(4.16)
m=1
(4.17)
(4.18)
This result is just one example of how a Kronecker delta may be removed from
an expression whenever one of its indices is a dummy sum index.
* Its important to inform your readers when you wish to temporarily suspend the summation conventions as we have done here. Some writers indicate that they do not wish for an index to be summed
by marking it with an overbar, as in A km mi , or by putting the index in parentheses, as in
A k ( m ) ( m )i or by typesetting non-summed indices with capital letters, as in A kM Mj .
28
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(4.19)
y ik
---------z i
(4.20)
i=1
BEWARE: avoid implicit sums as independent variables. A word of caution is essential here. When you employ the summation convention in derivatives, you should be extra careful to be clear about what the independent
variable is. Consider for example
w = b 11 + b 22 + b 33
(4.21)
w
w
------------ = --------------------------------------------- = 1
b kk
( b 11 + b 22 + b 33 )
(4.22)
w
w
w
w
------------ = ----------- + ----------- + ----------- = 1 + 1 + 1 = 3
b kk
b 11 b 22 b 33
(4.23)
or
The two answers are not the same, so we need to establish a precedence rule.
Our experience in reading the literature is that most authors intend for the
expression w b kk to be interpreted as Eq. (4.23). Thus, the precedence rule
is to always apply summations after taking derivatives. In other words, imagine that w is a function of nine b ij components. After finding all nine w b ij
derivatives, then w b kk is obtained by summing the three derivatives corresponding to i = j . To minimize confusion, we recommend that you write
w
kk , where ij = ---------b ij
(4.24)
cca B
ranno
(4.25)
29
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(4.26)
you would be wise to ignore the actual symbols used for the indices. Instead,
you should work to deduce the overall structure of the index placement. For
example, if you want to write down the formula for Y jim , then you could
always painstakingly convert Eq. (4.26) by replacing in all occurrences of i by
j , j by i , and k by m . However, doing it that way is error prone (especially
since you cannot do it step-by-step because you would violate the sum rules by
having four js in the expression after step 1 of replacing i by j ). People who
really understand indicial notation would construct the formula for Y jim by
spending a little time smelling the structure of the Eq. (4.26). If you look
carefully at that defining equation, you will note that the denominators in
the three terms on the right hand side have the same indices as those on the
left hand side and they also appear in the same order. Thus, your first step
to constructing the expression for Y jim would be to write a partial skeleton
formula as
a ?? a ?? a ??
Y jim = ---------+ ---------- ---------- ,
x j
x i x m
(4.27)
where the ?? stands for indices not yet inserted. Again looking at the structure of Eq. (4.26) you would note that the subscripts on each a ?? are simply
the other two indices not already in the denominator. Furthermore, those
subscripts are placed in an order that is a positive permutation of the free
indices moving clockwise from the index already placed in the denominator.
Specifically, the positive permutations of jim are: jim , imj , and mji .
Because the first term in Eq. (4.27) has x j in the denominator, we know that
the other two indices for the first term must be im . Similarly, the second
term has x i in the denominator, so the other two indices must be mj (not jm
because we need to select the same ordering as the positive permutation), etc.
Thus, the final expression is
30
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
a im a mj a ji
- + ------------ ---------Y jim = ----------x j
x i x m
(4.28)
(4.29)
To smell the structure of the index placement, you might find it useful to
mentally replace the indices with numbers representing the order in which
you should write them on the paper:
U 1234 = T 1324 + W 1423
(4.30)
Thus, for example, if you want to write the formula for U pqim , then you
should again start with a partially completed skeleton in which you place only
the first two indices:
U pq?? = T p?q? + W p?q?
(4.31)
Then you fill out the remaining two indices im by placing them in that order
on the first term and in the reverse order in the second term to obtain
U pqim = T piqm + W pmqi
(4.32)
(4.33)
cca B
ranno
(4.34)
31
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
would have to be written in traditional form by explicitly showing the summation sign:
3
k p k ( p k a )
(4.36)
k=1
(4.37)
(no sum on i ).
The phrase no sum on i tells the reader that the author wishes the index i to
be interpreted as a free index even though it appears exactly twice on the
right-hand side.
In tensor analysis, exceptions to the summation convention are rare, so it
is a very convenient notational tool, especially when an expression contains
numerous implied summations, as was the case in Eq. (3.50).
{a} = [ A]{v}
{b} = [ B]{w}
s = ai bi
s = { a }T { b }
(4.38)
The expressions on the right show the equivalent matrix expression for the
operations. Note, in the last equation, that a transpose of { a } is required in
the matrix equation. There is no need for a transpose on a i in the indicial
expression a iT is meaningless.
In the first two equations, the dummy summation index is j and the free
index is i ; hence, those equations actually represent three separate equations
for each value of the free index i . In the last expression, the dummy summation index is i , and there are no free indices (indicating that the equation is
just a single equation for a single scalar).
32
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
s = A ik v k B ij w j
(4.39)
which does not violate the summation convention. Written out in traditional
form, this equation may be written
3
s =
A ik v k B ij w j
(4.40)
i = 1k = 1 j = 1
cca B
ranno
(4.41)
Note that the Kronecker delta ij acts in a way that appears to have
changed the dummy summation index j on v j to become an i on v i with the
ij removed. This property holds because, in the summation over j , only one
term is nonzero (the one where i= j ).
This index-changing property holds in general. For example, the expression A ij v k ip can be simplified by noting that the subscript i on ip is a
dummy summation subscript. Therefore, ip may be removed if the other
occurrence of i is changed to a p . The simplified expression is therefore
A pj v k .
We have already used the index-changing property when we simplified
Eq. (3.31). The index-changing property applies in more complicated expressions involving multiple Kronecker deltas. Consider, for example, the expression T qms pk v q km id . Here the subscript k appears exactly twice, so it is a
dummy summation index, and the pk may be removed from the expression if
the other occurrence of k is changed to a p . This gives T qms v q pm id . Now we
33
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
note that the index m is a dummy sum index, so pm may be removed if the
other occurrence of m is changed to a p to give T qps v q id . The id in this
final expression may not be removed because both i and d are free indices,
not summation indices.
Now consider the expression mk A km . Noting that m is a summation
index, this may be simplified by removing the Kronecker delta mk if the other
occurrence of m is replaced by k to give A kk , which means A 11 + A 22 + A 33
and which is therefore equivalent to the trace operation of Eq. (3.15).
(4.42)
Viewed differently, this merely says that the trace of the 3 3 identity matrix
is equal to 3.
34
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
and w is
(5.1)
v w vk wk
Written out in full,
v w = v1 w1 + v2 w2 + v3 w3
In matrix notation, this can be written
v1
(v w)
= v2
v3
(5.2)
w1
w2 = { v } T { w }
w3
w1
= v 1 v 2 v 3 w 2 = <v> { w }
(5.3)
w3
The magnitude of a vector v is given by
2
2
v = v = + v 1 + v 2 + v 32 = + v v
Geometrically, the dot product can be written
(5.4)
v w = vw cos vw ,
(5.5)
v
where v and w are the magnitudes of v and w , respec-
tively, and vw is the angle between v and w. The dot
product is commutative:
vw = wv
It also satisfies the inner product positivity rule
cca B
ranno
vw
(5.6)
35
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
v v > 0 if v 0
v v = 0 if and only if v = 0
(5.7)
ij = 1 if i = j
0 if i j
(5.8)
(5.9)
v = v1 e1 + v2 e2 + v3 e3
(5.10)
Dotting both sides by e i , invoking Eq. (5.8), and finally using the index-chang
ing property of the Kronecker
delta gives
v e i = v k e k e i = v k ki = v i
(5.13)
which is equivalent to Eq. (5.11). This method for finding the component of a
vector might seem at first glance to be trivial and of no obvious use. Suppose
however, that { e 1*, e 2*, e 3* } is a different basis. Further suppose that we do
v with respect to the original (unstarred) basis, but
of
know the components
we wish to determine the components of this vector with respect to this new
(starred) basis. Then v*i = v e*i . As a specific example, suppose that
* Keep in mind that we are considering only ordinary engineering vectors having real components. If
complex components were allowed, the inner product would be written v w v k w k , where the
overbar denotes the complex conjugate.
36
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
1
--- ( 3e
1
5
1
--- ( 4 e
1
5
e 1* =
+ 4e 2 ) , e 2* = e*3 , and e 3* =
+ 3e 2 )
(5.14)
Suppose that we know that, with respect to the unstarred basis the vector v is
given by
v = 2e 1 + 5e 2
(5.15)
Then the components with respect to the unstarred basis are just v 1 = 2 ,
v 2 = 5 and v 3 = 0 , whereas the components of this vector with respect to the
starred basis are
6 20
26
1
v 1* = v e 1* = ( 2e 1 + 5e 2 ) --- ( 3e 1 + 4e 2 ) = --- + ------ = -----5
5 5
5
(5.16)
v 2* = v e 2* = ( 2e 1 + 5e 2 ) ( e*3 ) = 0
(5.17)
8 15
7
1
v 3* = v e 3* = ( 2e 1 + 5e 2 ) --- ( 4 e 1 + 3e 2 ) = ------ + ------ = --5
5
5
5
(5.18)
The method presented in this section works only when the new basis is
orthonormal; see page 51 to learn how to find the components of a vector with
respect to an irregular (non-normalized and/or non-orthogonal) basis.
part f o ( v ) as
where
1
f e ( v ) --- [ f ( v ) + f ( v ) ]
2
1
f o ( v ) --- [ f ( v ) f ( v ) ]
2
(5.20)
Homogeneous functions
A function f is said to be homogenous of degree k if it satisfies
f ( v ) = k f ( v ) for all positive scalars .
(5.21)
cca B
ranno
(5.22)
37
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
That is,
degree k1.
f ( v ) = k 1 f ( v )
(5.23)
w v
(a)
vw
v2
e2
e1
v1
(b)
(c)
Figure 5.1. Finding components via projections. (a) scalar component of v in the direction of w , (b) scalar component of w in the direction of v , and (c) the scalar components of
The operation
w
---------where w
(5.24)
that
is irrelevant thats why the formula depends only
on the unit vector w
points in the direction of w . The formula of Eq. (5.24) would not change if we
positive scalar . The result changes sign if w is
were to multiply w by any
replaced by w .
,
vw
38
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
the direction
v
(5.25)
where v -------w v ,
v
Cross product
The cross product between two vectors, a and b
is perpendicular to both a and
The unit vector n
defined by the right hand
rule
b , with a direction
when
sweeping from a to b through the angle
used.
The parallelogram in the illustration has an
orientation perpendicular to a b . The area of
the parallelogram equals the magnitude
of a b .
If u = a b , then the components of u are
u1 = a2 b3 a3 b2
u2 = a3 b1 a1 b3
u3 = a1 b2 a2 b1
ab
a
Figure 5.2.
Cross product
(5.27)
Hence,
(5.28)
a b = ( a 2 b 3 a 3 b 2 )e 1 + ( a 3 b 1 a 1 b 3 )e 2 + ( a 1 b 2 a 2 b 1 )e 3
Heuristically, this equation is often written as the determinant of a matrix:
ab
e1 e2 e3
det a 1 a 2 a 3
cca B
ranno
(5.29)
b1 b2 b3
39
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
e1 e1 = 0
e2 e1 = e3
e3 e1 = e2
e1 e2 = e3
e2 e2 = 0
e3 e2 = e1
e1 e3 = e2
e2 e3 = e1
e3 e3 = 0
(5.30)
(5.31)
(5.32)
In the last step, we applied Eq. (5.31). Noting that the final result is the sum
over k of an expression times e k , the k th component of the cross product oper
ation must therefore be
( a b ) k = a i b j ijk
(5.33)
This formula relies on our previously stated assumption that all component formulas in this report are referenced to an orthonormal right-handed
basis. If you use an orthonormal left-handed basis, then the above formula
would be ( a b ) k = a i b j ijk . Some authors take this situation as an implica
tion that the permutation symbol for left-handed base vectors should be
defined as the negative of the right-handed definition. This is misleading and
wrong-headed. Whats really going on is that the cross product operation can
be most generally written as a special third-order alternating tensor operat
ing on the vectors a and b . [Namely, using notation to be defined later
in this
ijk
ijk = ijk if the basis is orthonormal but left-handed. The ijk components
have yet a different form if the basis is non-orthogonal or non-normalized.*
Note that we keep the definition of the permutation symbol unchanged only
the components of change form upon a change of basis. We reiterate that,
40
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
of the basis, the final result for a b is the same in all cases. In other words,
a b represents a particular and unique vector that does not depend on the
underlying basis. Only the method that you must use to determine this unique
vector varies depending on the choice of basis.
[ a, b, c ] ( a b ) c
(5.34)
It can be shown that the triple scalar product has the same value if the vectors
are positively permuted and it changes sign if the vectors are negatively permuted. Specifically,
and
[ a, b, c ] = [ b, c, a ] = [ c, a, b ]
cca B
ranno
(5.35)
[ a, b, c ] = [ c, b, a ] = [ b, a, c ] = [ a, c, b ]
(5.36)
Physically, the absolute value of [ a, b, c ] equals the volume of the parallelepi of [ a, b, c ] will be positive if the vectors
ped formed by a , b , and c . The sign
meaning that when
the thumb of the right hand
form a right-handed
triad,
points in the direction of a and the index finger points in the direction of b
then the middle finger will point roughly in the direction of c (i.e., if the mid with c ). If the
dle finger were a vector, it would have a positive dot product
sign of [ a, b, c ] is negative, then the triad is said to be left-handed.
( e i e j ) e m = ijk km = ijm
(5.37)
where ijk is the permutation symbol defined in Eq. (3.21). Expressing the
above result with the free index m replaced by k gives
ijk = [ e i, e j, e k ]
(5.38)
The triple scalar product [ e i, e j, e k ] is certainly well-defined if the basis is
is
[ e , e , e ] = . As mentioned earlier, this
left-handed; in fact, the result
ijk
i j for
k redefining
fact should not be used as justification
the permutation symbol
when using a left-handed basis. We recommend always defining the permutation symbol such that 123 =+1 even if the basis is left-handed.
Even though the term tensor has not yet been defined, its worth mentioning here (for
future reference) that there is a straightforward formula for constructing a third-order permutation tensor in terms of any basis including left-handed and even non-orthonormal. The
permutation tensor components with respect to a non-orthonormal basis take yet a different
41
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
form (covered in the more advanced Ref. [4]), and the process of computing the triple scalar
product becomes more complicated. Nonetheless, the basic concept is the same: the triple scalar product of three vectors equals the triple inner product of the alternating tensor into the
dyadic multiplication (defined later) of the three vectors. It just happens that the components of the alternating tensor equal the permutation symbol when the basis is orthonormal
and right-handed. It is for this reason that we denote the alternating tensor by a symbol ijk
that differs from its right-handed orthonormal components ijk . A similar notational choice
is also made when we identity tensor, I , by a symbol that differs from its right-handed
Axial vectors
Even though we have not yet covered the meaning of the word tensor,
this is still the most appropriate place (for future reference, once the meaning
of the word tensor is learned) to discuss axial vectors and axial tensors.
Given any vector a , we can always construct a skew-symmetric axial tensor
defined by
A a
(5.39)
We have not yet defined what is meant by the word tensor, nor have we
defined what is meant by the above tensor notation. For now, A should be
(5.40)
Expanding out the implied summation over the index k (noting that all terms
for which i= j are zero, the matrix for the axial tensor is related to the components of the vector a according to
0 a3 a2
[ A] =
a3
0 a1
a2 a1
(5.41)
Conversely, given any tensor B , the associated axial vector can be constructed
by
1
a = --- : B
2
(5.42)
(5.43)
(5.44)
42
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
1
a 2 = --- ( B 13 B 31 )
2
(5.45)
1
a 3 = --- ( B 21 B 12 )
2
If A is the skew-symmetric part of B , then
a 1 = A 32
cca B
ranno
(5.46)
(5.47)
a 2 = A 13
(5.48)
a 3 = A 21
(5.49)
That is,
(b c) = bc cb
(5.50)
We placed this result here because of its intimate connection with cross products. Expressions like this show up frequently in single crystal plasticity.
43
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
by w
itself is called the orthogonal vector projection of v onto
direction of w
denote
this vector by v Dw (which may be read as the part
of v
w , and we will
the direction of w ). The mathematical
in
definition is
w
Dw
)w
---------where w
v
= (v w
(6.1)
For some applications, it is prudent to recognize that the projection operation involves two vectors, v and w , so it is sometimes advisable to instead
define a binary* operator D such that D ( v, w ) denotes the part of v in the
vw
(6.2)
D ( v, w ) = ------------- ww
Clearly,
D ( v, w ) D ( w, v )
(6.3)
* The term binary is merely a fancy way of saying that the operator has two arguments.
44
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
v Dw
denote by v Pw or by
vector v that is perpendicular
to w , which we will
P ( v, w ). It is defined by
)w
v Pw = v ( v w
(6.4)
or
vw
(6.5)
P ( v, w ) = v -------------
ww
Note that v Pw lies in the plane perpendicular to w . Also note that
P ( v, w ) + D ( v, w ) = v
(6.6)
This equation shows more clearly that the operators P and D decompose the
vector v into two parts, one parallel to w and the remainder perpendicular
to w .
lems, draw a plane perpendicular to w ) that passes through the tip of the vector v . These two lines (or, for 3D problems, the line and the plane) will
intersect
at a single point, which we will call A . Then v Dw is the vector
(b)
(a)
tip
(c)
A
v Pw
w Dv
A
tail
v Dw
to
lines
parallel
w Pv
lines
parallel
to v
tail
tip
tip
v De2
e2
tail
e1
v De1
by the formulas
and
w Dv = D ( w, v ) = ( w v )v
Pv
w
= P ( w, v ) = w ( w v )v
v
where v -------v
cca B
ranno
(6.7)
45
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Figure clearly shows that this decomposition of w results in completely different vectors as the decomposition of v .
The last sketch in Fig. shows the decomposition of a vector v into parts
that are aligned with the base vectors. In particular, note that we labeled the
vertical part of the vector as v De2 because, in three dimensions v De2 v Pe1 .
(6.8)
v De2 = 3e 2
whereas
v Pe1 = 3e 2 + 7e 3 .
(6.9)
direction of a second vector b . Since this operation involves two vectors, you
adjective orthogonal is used to indicate that each point on the vector a maps
a family of lines that are orthogonal to b . The set of all vectors parallel to b
forms a one-dimensional subspace because any member of the set may be written in the form b , which involves only one arbitrary scalar .
If n is a unit vector, then any vector x can be projected into a part that is
to n and simultaneously perpendicular
parallel
to the plane whose normal is
(6.10)
p = n(n x)
In other words, p is just x Dn . This projection has rank 1 because the target
line parallel to p . If one were to compute the
tor p .
46
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Plane perpendicular to
(b)
(a)
Plane perpendicular to
Figure 6.2. (a) Rank-1 orthogonal projection, and (b) Rank-2 orthogonal projection.
The dashed lines show the projection direction. For the Rank-1 orthogonal projection, the
dashed lines point from a location on x to a target line that is parallel to the unit vector n
q = x n(n x) .
(6.11)
would
be the 1-2 plane. Substituting
n=e 3 into Eqs. (6.10) and (6.11) and
applying Eq. (5.11) gives
p = e3 ( e3 x ) = x3 e3 .
q = x x3 e3 = x1 e1 + x2 e2 .
cca B
ranno
(6.12)
(6.13)
47
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Thus, the projected vector p is the part of x that is in the 3-direction, and q
many engineering applications,
the plane.
plane. This
time, however, we want
to be aligned with a different vector a . Referring to Fig. 6.3, we can see that
(6.14)
x = q + a ,
To find the value of , we impose the condition that the vector q must lie in
(6.15)
b x = 0 + (b a) .
Solving for and substituting the result into (6.14) gives the desired formula
for the oblique projection:
48
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
a(b x)
Q ( x ) = x --------------------(6.16)
- .
ab
Naturally, Eq. (6.11) is a special case of the more general Eq. (6.16), obtained
by choosing a =b =n . In other words, the projection is orthogonal only if a is
b
.
proportional to
q = Q ( x ) , where
a(b x)
p = P ( x ) , where P ( x ) = -------------------- - . (6.17)
ab
The result of this operation will always be a
vector that is simply a scalar multiple of a .
cca B
ranno
Plane perpendicular to
Note that the operator P has the prop- Figure 6.4. Rank-1 oblique proThe path obliquely intererty that P ( P ( x ) ) = P ( x ) . Physically, this jection.
sects
the
plane.
a vector that is
means that projecting
already in the target space will just give you back that vector unchanged (your
shadow has no shadow other than itself). This property is, in fact, used to
define the term projection. The operator Q has the similar property that
Q ( Q ( x ) ) = Q ( x ) , so it is therefore a projector.
Complementary projectors
The operator P is called the complement of Q . In general, two operators P
and Q are called complementary projectors if the following properties hold
P( P( x)) = P( x)
(6.18)
(6.19)
Q(Q( x)) = Q( x)
(6.20)
Q( P( x)) = P(Q( x)) = 0
P( x) + Q( x) = x
(6.21)
The last property states that the sum of P and Q must be the identity operation (this is an abstract way of saying that the P and Q operators permit an
additive decomposition of the vector x ). The rank of P plus the rank of Q
must sum to equal the dimension of the space, which is 3 for ordinary physical
vectors.
49
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
tions matter.
Because the magnitudes of a and b are arbitrary, it is convenient in cer with a
and b , which have the same orientatain applications to replace them
satisfy
b =1
a
(6.22)
One possible way to construct a and b is
a
b
= ---------------a
b = ---------------(6.23)
and
ab
ab
and b are called dual vectors.
When the property of Eq. (6.22) holds, a
The component form of Eq. (6.16) take a particularly simple form when
expressed using a nonorthogonal basis having the covariant* base vector
g 3 = a and the contravariant base vector g 3 = b . Namely
1
2
q = x g1 + x g2 ,
(6.25)
This result is strongly analogous to Eq. (6.13), with the key difference being
that the g k base vectors are nonorthogonal and not generally of unit length.
The projector
Q is a rank-2 projector because its range space is two dimensional.
The null space of the projector Q is the set of all x for which Q ( x ) = 0 .
delta of Eq. (5.9). Thus, for example, g 1 is perpendicular to the plane formed by g 2 and g 3 . A
called covariant and superscripted quantities are contravariant. (A cute mnemonic is cogo-below). Non-orthonormal tensor analysis is reviewed in Refs. [5,4].
50
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
two dimensional. Since P and Q are complementary projectors, the null space
of one is the range space of the other! Later on, projectors will be defined for
higher-dimensional spaces and the null spaces will be more interesting.
Figure (6.5) shows two vectors, x and y ,
cca B
ranno
Q ( x )=Q ( y )
The brute force simplest way to determine values of the scalar coefficients is to
simply dot both sides of this equation by a to obtain
x a = (a a) + (b a) + (c a)
(6.29)
We can similarly dot both sides of Eq. (6.28) by b and then separately by c to
(6.29) forms a set of three
The right-hand side of Fig. 6.1 showed how an arbitrary vector x may be
decomposed into the sum of parts that are aligned with the orthonormal
{ e 1, e 2, e 3 } basis. However, a basis does not have to be orthonormal any
linearly
three
independent vectors { g 1, g 2, g 3 } may be alternatively used as a
51
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
P3 ( x )
g2
g3
P2 ( x )
angles 90
g1
P1 ( x )
Figure 6.6. Three oblique projections. The indicated plane contains g 1 and g 2 ,
to
while g 3 forms an oblique angle to the plane. The dual vector g 3 is proportional
In other words, g 1 must be perpendicular to g 2 and g 3 , so it must be propor ) . The constant of proportionality
be given by
must
1
1
1
g 1 = ------ ( g 2 g 3 ) , g 2 = ------ ( g 3 g 1 ) , g 3 = ------ ( g 1 g 2 )
(6.32)
Jo
Jo
Jo
The three P k operators are complementary projections because they satisfy
the following properties:
P i ( x ) if i= j
Pi ( P j ( x ) ) =
if i j
0
(6.34)
P1 ( x ) + P2 ( x ) + P3 ( x ) = x
(6.35)
and
52
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
Generalized projections
A transformation P is called a projection (or idempotent) if and only if
P( P( x)) = P( x) ,
(6.36)
which merely states that projecting a vector that has already been projected
will give you the same projected vector right back. Consider, for example, the
function of Eq. (6.16):
a(b x)
P ( x ) = x --------------------(6.37)
ab
Physically, we know that this function is a projector. To prove it rigorously, we
operate on both sides of Eq. (6.37) by P to obtain
a(b x)
a b x -------------------- -
ab
a(b x)
P ( P ( x ) ) = x --------------------(6.38)
- -------------------------------------------------------ab
ab
Simplification shows that the right hand side reduces back to P ( x ) . Hence,
Linear projections
A projection is linear if and only if
(i) P ( x ) = P ( x ) , for all , and
The projection of Eq. (6.37) is easily seen to be linear even though it is oblique.
Likewise, Eq. (6.33) is linear. Physically, a projection is linear if one can take
the projection of a linear combination of vectors and the result is the same as
first projecting the vectors and then taking their linear combination.
Nonlinear projections
An example of a nonlinear projection is
x
P ( x ) = ---------------(or P ( x )=0 if x =0 )
(6.40)
xx
This function is indeed a projection because Eq. (6.36) holds. The first linearity test (i) in Eq. (6.39) fails because, for example, P ( 2 x ) 2P ( x ) , so this pro
a vector to a
jection is nonlinear. Geometrically, the above operator transforms
unit vector in the same direction. Hence, the range set for this transformation
consists of vectors pointing from the origin to a point on the unit sphere. The
transformation of Eq. (6.40) is non-linear because the range (the target surface of the sphere) is geometrically curvilinear. The level lines (i.e., the light
rays) are straight lines that emanate radially from the origin.
53
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Nonlinear projectors can also be constructed for which the range is a linear
manifold (i.e., a line or a plane) but the path to this range space is curvilinear.
For example, the transformation
[ x a ( x b ) 2 ]a
P ( x ) = -----------------------------------------------(6.41)
a (a b)
is a projection because P ( P ( x ) ) = P ( x ) . It has a linear range space (vectors
parallel to a ), but the projection is nonlinear because the paths to the range
space (i.e., the level lines) are curvilinear. The light rays are bending as they
approach the target space.
A very important nonlinear projection transformation from the field of continuum mechanics is the polar decomposition (see page 136).
Self-adjoint projections
A projection P is self-adjoint if and only if
u P ( v ) = v P ( u ) for all u and v .
(6.42)
a(b v)
a(b u)
u v -------------------- = v u -------------------- ab
ab
Simplification gives
(6.43)
54
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
decomposed uniquely as
xP
xQ
x = x P + xQ
(6.45)
such that x P is in the range space of the projector P and x Q is in the null satisfy
space of the projector P . By this we mean that x P and x Q will
P( xP) = xP
(6.46)
P( xQ ) = 0
(6.47)
xP = P( x)
(6.48)
(6.49)
xQ = Q( x ) ,
where
Q( x) = x P( x)
(6.50)
The function Q is itself a projector. Some key properties of these complementary projectors are
P( P( x)) = P( x)
Q(Q( x)) = Q( x)
P( x) + Q( x) = x
cca B
ranno
(6.51)
* For example, the principle of conservation of mass merely states that what goes in must come out
or stay there. The mathematical representation of this statement is less obvious.
55
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
complementary projectors.
The transformation P (x ) is an orthogonal projection if and only if P is self-adjoint, in which case x P x Q will also be zero.
56
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
* This is true for some practical purposes, but not for all practical purposes.
Often, determining if one definition implies the other is a difficult task. The person who defines a
tensor according to basis transformation rule is actually defining a particular class of tensors,
whereas the definition in terms of linear transformations has broader abstract applicability, but
becomes bogged down in the more subtle question: what is a vector? Answering this question for
engineering mechanics applications eventually also comes around to the need to introduce basis
transformation criteria, so in this arena, both definitions are equivalent. See page 146.
57
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
result
(i)
for all scalars
f ( v ) = f ( v )
and (ii)
for all vectors u and v
f (u + v) = f (u) + f (v)
(7.1)
Importantly, these three vectors can be computed once and saved for all time.
Usually these vectors are stored as columns of a 3 3 matrix [ F ] so that
[ F ] = [{ f 1}{ f 2}{ f 3}]
(7.3)
F ij = e i f j
f j = F ij e i
(7.4)
x = x1 e1 + x2 e2 + x3 e3
(7.5)
Operating on this vector by the linear function f and applying the linearity
property, shows that the action of the linear transformation on an arbitrary
vector v can be rapidly computed by
f ( x ) = x1 f ( e1 ) + x2 f ( e2 ) + x3 f ( e3 )
= x1 f 1 + x2 f 2 + x3 f 3
(7.6)
(7.7)
58
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
f ( x ) = x j ( F ij e i ) = ( F ij x j )e i
(7.8)
y = f ( x ) = yi e .
(7.9)
(7.10)
The matrix [ F ] is called the matrix of components of the tensor associated with the linear transformation f . For any linear function, there exists an
associated tensor, and vice versa. Consequently, many authors [e.g., 19] define
a tensor to be any linear transformation of vectors to vectors.* We prefer to
keep the two ideas separate in our presentation. After all, we frequently know
the linear transformation before we know the tensor itself. For example,
when b is a known constant vector, the operation f ( x ) = b x is linear with
operation
is not immerespect to x . However, the tensor associated with this
diately obvious. The operation is linear, so we know that a tensor (representable by a matrix) must exist, but what is the tensor matrix that corresponds to
f ( x) = b x ?
It is critical to recognize that the components of the matrix F ij are, by construction, referenced your chosen basis. Thus, implicitly, a tensor must consist
of both a matrix of components and an associated basis. If the basis changes,
then the component matrix changes in a specific way (explained later), which
is why some people [e.g., 18] define a tensor to be a set of components that
transform in the necessary way upon a change of basis.
The intimate dependence of the F ij components on the underlying basis
{ e 1, e 2, e 3 } is well emphasized by using the following basis notation for ten
sors:
F =
+
+
F 11 ( e 1 e 1 ) + F 12 ( e 1 e 2 ) + F 13 ( e 1 e 3 )
F 21 ( e 2 e 1 ) + F 22 ( e 2 e 2 ) + F 23 ( e 2 e 3 )
F 31 ( e 3 e 1 ) + F 32 ( e 3 e 2 ) + F 33 ( e 3 e 3 )
cca B
ranno
(7.11)
* As clarified by Simmonds, ...To say that we are given a 2nd order tensor T means that we are told
Ts action on (i.e., where it sends) any vector v. Thus two 2nd order tensors S and T are said to be
equal if their action on all vectors is the same...
The distinction is analogous to the unique correspondence between animals and their DNA
sequences. Disregarding cloning and identical twins, there is exactly one animal for each DNA
sequence and vice versa, but this does not mean that animals and DNA sequences are the same
thing. Likewise, in tensor analysis, one often has a well-characterized linear transformation without
having an explicit expression for the associated tensor (even though we know it exists).
59
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
The dyadic multiplication symbol is defined below. For now, the presence
of e i e j next to F ij should be regarded as a way to indicate to the reader
F = F ij ( e i e j )
(7.12)
We will usually omit the symbol for dyadic multiplication so that two vectors written side by side are understood to be multiplied dyadically, and the
above equations would be written more compactly as
F =
F 11 e 1 e 1 + F 12 e 1 e 2 + F 13 e 1 e 3
F 21 e 2 e 1 + F 22 e 2 e 2 + F 23 e 2 e 3
F 31 e 3 e 1 + F 32 e 3 e 2 + F 33 e 3 e 3
+
+
(7.13)
and
F = F ij e i e j
(7.14)
Fx
f ( x)
(7.15)
Most authors do not use our double underline (under-tilde) convention to indicate the order of a tensor, so you will typically see tensors typeset simply in
bold, F. Furthermore, many authors do not use the raised dot notation (hence,
to them, Fx means the same thing as what we write as F x . As will be fur
ther explained later, our notation lends itself better to heuristic self-explanatory interpretations of the intended meanings of operations, which becomes
indispensable when working with higher order tensors in modern material
modeling applications.
Recall that many people define a tensor to be a linear operation taking vectors to vectors. The basis dyad e i e j is itself a tensor, and it has a component
matrix that contains all zeros except for a 1 at the ij position. Thus, the
matrix form of Eq. (7.13) is simply
60
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
F 11 F 12 F 13
F 21 F 22 F 23
100
010
001
F 11 0 0 0 + F 12 0 0 0 + F 13 0 0 0
000
000
000
F 31 F 32 F 33
+
F 21
000
000
000
1 0 0 + F 22 0 1 0 + F 23 0 0 1
000
000
000
F 31
000
000
000
0 0 0 + F 32 0 0 0 + F 33 0 0 0
100
010
001
(7.16)
(7.17)
(7.18)
takes on meaning
Note that the right hand side of Eq. (7.19) equals the vector a times a sca (b a) v
lar, ( b v ) . Thus, ( a b ) v is proportional to a . Similarly,
That is,
abba
in general
cca B
ranno
(7.21)
61
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
is equivalent to
tional to b . The condition on the proportionality constant
r and/or s is zero.
The above discussion reiterates that dyadic multiplication does not commute; i.e.,
abba
(7.24)
Referring to Eq. (7.23), we obtain the necessary and sufficient condition for
a b to commute:
(7.25)
a b = b a if and only if b = a for some .
In other words, the two vectors in the dyad would have to be multiples of each
other in order for a dyadic multiplication to commute.
In addition to being a more compact notation, this convention for dyadic multiplication also has the appealing property that the definition of Eq. (7.19) can
be written
(ab) v = a(b v) ,
(7.27)
62
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
allowing us to drop the parentheses without ambiguity.* We advocate retaining the explicit symbol only when it is needed for clarity.
Dyads are like defensive linemen in football essential, but rarely of
much interest per se. They are merely abstract mathematical objects, which
(up to this point) have no apparent practical use. After we define addition and
scalar multiplication for dyads, we will see that tensors (which are of great
practical importance) are always expressible as a linear combination of the
nine possible dyads between the orthonormal base vectors. These nine basis
dyads, { e 1 e 1, e 1 e 2, , e 3 e 3 } , will form a basis for tensors just as the three
{e , e , e } form a basis for ordinary vectors. With the no symbase vectors
2 3
1 (7.12)
may be written
bol notation, Eq.
F = F ij e i e j ,
(7.28)
In this expression, the indices i and j are summed from 1 to 3. Thus, the
above expression is a linear combination of dyads, the meaning of which is
described in more detail below.
(7.29)
We can define a 3 3 matrix whose ij components are a i b j . Then the expression in Eq. (7.29) can be written
[ a i b j ]v j
(7.30)
cca B
ranno
(7.31)
which is the matrix representation of the left-hand side of Eq. (7.27). Using
conventional matrix notation, note that
* Some people strenuously object to this side-by-side notation for dyadic multiplication. They argue
in favor of the symbol because dyadic multiplication is different from scalar multiplication.
By the same logic, however, the identity ( + )v = v + v should also be notationally objection
able because addition between vectors is different from addition between scalars. Likewise, the
notation dy dx would be objectionable because derivatives are not really fractions. In both mathematics and engineering, we routinely overload operations for good reason: heuristic notational
advantages. The meanings of the overloaded operations are implied by the nature of the arguments.
63
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
a1 b1 a1 b2 a1 b3
a1 b1
[ a b ] = a2 b1 a2 b2 a2 b3 = a2 b2
a3 b1 a3 b2 a3 b3
a3 b3
= { a }<b>
(7.32)
The result is a square matrix. The reader should contrast this operation with
the similar-looking scalar-valued operation of Eq. (5.3) in which the transpose
merely occurs in a different place!
The sum of two or more dyads takes on meaning only when operating on a vector. The sum of dyads also has a matrix interpretation. Specifically, the matrix
[ a b + c d ] associated with the sum of two dyads is obtained by simply adding
matrices
the
for [ a b ] and [ c d ] .
The
sum of three dyads is also a distinct matha distinct new abstract
object.
ematical abstraction that in general cannot be written as either a single dyad
or the sum of two dyads. We will later demonstrate that the sum of four or
more dyads can always be reduced to the sum of three or fewer dyads (provided that the vectors belong to ordinary 3D space). Thus, the sum of four or
more dyads is not a new object.
for all v
[ (ab)] v = a(b v)
(7.36)
64
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Using the dyad symbol for clarity, we note that the scalar multiple may be
regarded as an external multiplier or it may be absorbed into any one of the
vectors in the dyad:
(7.37)
(a b) = ( a) b = (a ) b = a ( b) = a (b )
If the scalar is placed on the right side of the dyad or between the two vectors,
it can be moved to the left side. In other words,
(7.38)
(ab + cd + e f ) = ( ab + cd + e f )
cca B
ranno
(7.39)
( a b + c d + e f + g h ) v = a ( b v ) + c ( d v ) + e ( f v ) + g ( h v ) (7.40)
Throughout this report, our vectors are ordinary vectors in 3D physical
space. Consequently, any set of four vectors must be linearly dependent. Thus,
at least one of the vectors in the set, { b, d, f , h } , can be written as a linear
for illustration
h = 1 b + 2 d + 3 f
(7.41)
The left-hand side is the sum of four dyads. The right hand side is the sum of
three dyads. This proves that any sum of four or more dyads can always be
reduced to three or fewer dyads. Consequently, the sum of four or more dyads
is not a new abstract object! The sum of any number of dyads is generally
referred to as a dyadic, but we will use the term tensor.
65
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
linear
associated with this
transformation. Dyads are the most rudimentary
tensors. As emphasized earlier, the sum of two dyads generally cannot be
reduced to a single dyad. Likewise, the sum of three dyads cannot be reduced
to the sum of fewer dyads. However, the sum of four or more dyads can always
be reduced to the sum three or fewer dyads. Consequently, a second-order tensor may be defined as any sum of dyads. The term dyadic is also used to
mean the same thing. This definition is equivalent to the more traditional definition of a tensor in terms of linear transformations from vectors to vectors in
the sense that one definition implies the other. The sum of dyads definition
is more useful in certain settings, especially when considering a tensor as a
higher-dimensional vector.
For clarity, we typeset second-order tensors in bold using two under-tildes.
Thus, A , B , U would denote second-order tensors. Tensors are extremely
useful in physical applications. For example, tensors represent stress, strain,
inertia, electrical permittivity, and many other important physical quantities.
Once a physical quantity is proved to be representable by a tensor, a wealth of
theorems from abstract tensor analysis then apply to the tensor, furthering its
physical interpretation.
66
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
tensors form a three dimensional space; however the three dyads are not the
same for all tensors in fact, they are not even unique for any given tensor.
To determine the dimension of the space of tensors, we need to determine how
many scalars are needed to uniquely define a tensor.
Given a tensor A (i.e., given a sum of dyads), we know that there exist*
Each of the vectors may be expanded in terms of a basis ( a = a i e i , etc.), in
(7.45)
(7.46)
a b = ai b j ei e j
Performing similar expansions for the other terms in Eq. (7.44), we see
that any sum of dyads (i.e. any tensor) can be written as a linear combination
of the nine possible basis dyads e i e j for i and j ranging from 1 to 3. That is, for
A = A ij e i e j
(7.47)
where there are an implied sums of i and j ranging from 1 to 3 for a total of
nine terms. The principal advantage of the representation in Eq. (7.47) is that
the A ij components are unique for a given orthonormal basis, just as the components of an ordinary vector are unique for a given basis.
A 21 A 22 A 23
(7.48)
A 31 A 32 A 33
A =
i=1j=1
A ij e i e j
(7.49)
cca B
ranno
(7.50)
* Keep in mind, we only need to assert that these vectors exist. In practice, the appropriate vectors are
almost never actually computed.
67
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A 11 A 12 A 13
[ A] =
A 21 A 22 A 23
A 31 A 32 A 33
(7.51)
ee
The e e subscript tells the reader that the ij component of the matrix is the
A 21 A 22 A 23
<lab>
(7.52)
A 31 A 32 A 33
would indicate to the reader that the components are referenced to the laboratory basis. Finally, the notation
A 11 A 12 A 13
[ A] =
A 21 A 22 A 23
< e 3 =n >
(7.53)
A 31 A 32 A 33
would tell the reader that the component matrix [ A ] is being displayed with
respect to a special purpose basis in which the 3-direction has be aligned with
a known unit vector n . This sort of display would be used when the compo
nent matrix is particularly simple with respect some particular (cleverly
selected) basis.
68
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
8. Tensor operations
Dotting a tensor from the right by a vector
Keep in mind that a tensor is a mathematical bookkeeping device, which
takes on physical meaning only when operating on a vector. The notation
A v represents the linear transformation associated with the tensor. Specifi
cally, applying the definition of a linear combination of dyads, the notation
A v must be interpreted as
A v = ( A ij e i e j ) v = ( A ij e i ) ( e j v ) = ( A ij e i )v j = ( A ij v j )e i
(8.1)
The ij component of
tion becomes
u i A ijT v j = v m A mn u n
cca B
ranno
(8.4)
We would like to change the dummy summation indices on the right hand side
so that we may compare it to the left hand side for arbitrary vectors u and v .
69
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
u i A ijT v j = v j A ji u i
(8.5)
Rearranging gives
u i ( A ijT A ji )v j = 0
(8.6)
The only way this can hold for all vectors u and v is if
A ijT = A ji
(8.7)
The above equation is often cited as the definition of the transpose, but the
direct notation definition of Eq. (8.3) is more general since it makes no
assumption that the underlying basis is Cartesian.
It is straightforward to show that
( A )T = AT
(8.8)
and
(8.9)
( A + B )T = AT + BT
( ab )T = ba
(8.10)
If a tensor is written in the form of Eq. (7.50), the transpose simply swaps
the base vectors. Specifically,
A T = ( A ij e i e j ) T = A ij ( e i e j ) T = A ij e j e i = A ji e i e j
(8.11)
change i to j
and j to i
swap
In writing the final form, we have merely emphasized that any symbol may be
used for the dummy subscripts. Namely, we may swap the symbols i and j
without loss. The ji component of A T is the coefficient of e j e i ; so the second
says that the ij
to-last expression says that A Tji = A ij . The final expression
component of A T is the coefficient of e i e j , namely A ji . Both statements are
Following an analysis similar to Eq. (8.1) we write
u A = u ( A ij e i e j ) = A ij ( u e i )e j = A ij ( u i )e j = ( u i A ij )e j
(8.12)
(8.13)
70
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Thus,
u A is a vector whose j th component is u i A ij .
Equivalently, changing the symbols for the indices,
(8.14)
T
uA = A u
(8.16)
(8.17)
(8.18)
cca B
ranno
(8.19)
(8.20)
A B = A ij B kl e i e j e k e l
Applying Eq. (5.8), we note that e j e k = jk so that
A B = A ij B kl e i jk e l
(8.21)
(8.22)
71
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(8.23)
A B = A ij B kl jk e i e l
Finally, imposing the index changing property, we may remove the jk Kronecker delta if all occurrences of j are replaced with k , giving
(8.24)
A B = A ik B kl e i e l
This result is a linear combination of A ik B kl times the dyad e i e l . Therefore,
B . Later
A B must represent a tensor whose il component is given by A
ik kl
on, the linear transformation operator associated with A B will be seen to
represent the composition of the linear transformations associated individually with A and B .
72
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
73
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(9.1)
74
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
1
d ---
x
dx dt
------------ = ---------------dt
x2
(9.2)
Tensor gradients are another good example. We can define direct notation
symbols for the derivative of scalars, vectors, or tensors with respect to other
scalars, vectors, or tensors as follows:
u i
du u i
du
ds s
------- --------- e k ,
--------------- ---------- e i e j ,
------------- e i e j e k , etc. (9.3)
d x x k
dw w j
d A A jk
By using a notation such as this, direct notation expressions for the chain rule
take forms that are very similar to what is already familiar for scalars. For
example, it can be shown that
du 1
dw
-------= -------(9.4)
dw
du
This formula looks like an ordinary application of the chain rule except that
the appropriate inner product is used between factors. The indicial form of the
above equation would be
F mn
s s a k s
s
-------- = ------- -------- + --------- --------- + -------------- ------------- x i a k x i F mn x i
x i
cca B
ranno
(9.6)
Basis notation is useful when working with more than one basis, or as an
intermediate notation when converting a direct notation expression to indicial
form. Indicial notation is often the most clear, though an alphabet soup of indices can be distracting when discussing conceptual issues. Furthermore, the
constant task of looking after the summation rules (changing j s to k s, etc.)
can be wearisome and error-prone. There are times when even indicial notation is vague. For example, what does the expression f ii mean? Is it
f ( tr ) or is it tr [ f ] ? The two are not the same. Likewise, does A ij1
75
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
are characteristic of potential flaws with indicial notation. The bottom line is
that all three notations are useful, and none is perfect. A competent analyst
should learn all notations and pick the one that is clearest for the application
at hand. A courteous analyst will always attempt to infer which notational
system their audience (or the majority of their audience) will be most comfortable with. Dont be one of those pompous idiots who insist that their preferred
system is the only way to go and therefore everyone else must have it
shoved down their throats.
and f ( x ) is linear in x
y i = F ij x j
(9.7)
(9.8)
y1
F 11 F 12 F 13
y 2 = F 21 F 22 F 23
F 31 F 32 F 33
y3
x1
x2
x3
(9.9)
matrix
is to recognize that the i th column of [ F ] contains the components of
the vector f ( e i ) .
For analytical applications, a direct notation form for the tensor is often
desired. The tensor F can be determined by simply differentiating the func
tion f ( x ) with respect to x .
76
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
if y = f ( x )
,
and
f
(
x
)
is
linear
in
x
yi
dy
where F = ------- = --------e i e j
(9.10)
x j
dx
An example of this statement is given in Eq. (11.23). Incidentally, the derivative ( ) x j denotes the partial derivative with respect to x j , holding the
other components of x constant. When applying Eq. (9.10) for finding F , it is
x i
dx
(9.11)
------- = --------e i e j = ij e i e j I
x j
dx
such that
I x = x for all x
(9.13)
matrix
100
[I] = 0 1 0
001
(9.14)
or, expanded out explicitly,
cca B
ranno
(9.15)
77
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
I = e1 e1 + e2 e2 + e3 e3 = ek ek
(9.16)
An alternative way to derive Eq. (9.14) is to recall that the i th column of the
tensor must contain f ( e i ) . Thus, the first column of the tensor must contain
the component array for f ( e 1 ) = e 1 . The component array for e 1 is
1
0 ,
0
(9.17)
so this must be the first column of the matrix. The second and third columns
are the component arrays for e 2 and e 3 , respectively. Thus we obtain the
such that
f ( x) = F x
(9.18)
that
g( x) = G x
(9.19)
Therefore, the composition of the two transformations must be given by
f ( g( x)) = F (G x)
(9.20)
The components of F G are
( F G ) ij = F ik G kj
(9.22)
Hence, the matrix for F G may be found by the simple matrix multiplica
tion, [ F ] [ G ] .
Eqs. (9.20) and (9.21) must both hold simultaneously for all x vectors.
(9.23)
78
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
the reader can heuristically write
(9.25)
F G = ( F ij e i e j ) ( G kl e k e l ) = F ij G kl ( e i e j ) ( e k e l )
(9.26)
Using Eq. (5.8), the user can note that e j e k = jk , so that the above equa
(9.27)
F G = F ik G kl e i e l
(9.28)
The remaining dyad is e i e l . Hence, the il components of F G must be
( F G ) il = F ik G kl
(9.29)
Aside from our choice of dummy and free indices, this result is identical
to (9.22). Throughout this report, we define our notation in such a manner
that heuristic analyses like these will always give the correct interpretation of
the notation. Thus, for example, the cross product of a tensor with a vector
would be interpreted
A v = ( A ij e i e j ) ( v k e k )
= ( A ij v k e i ) ( e j e k )
cca B
ranno
(9.30)
A v = A ij v k pjk e i e p
(9.31)
This final form contains a dyad e i e p . The reader (who might never have heard
would be able to conclude that A v must
of crossing a tensor into a vector)
be a second-order tensor, with ip components given by
( A v ) ip = A ij v k pjk
(9.32)
Similarly, a reader would be able to deduce the meaning of two tensors written
side-by-side with no symbol between them as
79
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A B = A ij B kl e i e j e k e l
(9.33)
We have already introduced the notion that any linear combination of e i s is a
Genvector, and any linear combination of e i e j dyads is a second-order tensor.
basis triads e e e which will be
eralizing this idea, we will later introduce
k
i j combination
identified with third order tensors. Likewise, any linear
of
e i e j e k e l must be a fourth-order tensor. Hence, the notation A B in Eq. (9.33)
represents
must
a fourth-order tensor whose ijkl components must be A ij B kl .
Recall that u v means the same thing as u v . Likewise, when publishing for
because your readers might confuse A B with tensor composition, which we
here denote as A B .
Now suppose that f is defined to undo the action of g so that
(9.34)
Then f is called the inverse of g , and is denoted g 1 . The tensor associated
with the inverse of g is denoted G 1 . Hence, putting Eq. (9.35) into (9.34)
(9.36)
G 1 G = I ,
(9.37)
which means that the matrix for G 1 is obtained by the inverse of the matrix
for G .
This definition might seem quite strange until you think about it physically.
The magnitude of the cross product u v equals the area of the parallelogram
of the cross product is perpendicular to
formed by u and v and the direction
of Eq. (9.38)
is the new
area vecG v . Thus, the vector on the right-hand-side
formed
tor
by the transformed vectors. Though not at all obvious that it should
80
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
be so, it turns out that this new area vector is linearly related to the old area
vector u v . Since the relationship between the old and new areas is linear,
exist a tensor, which we denote G C , that can act on the old area to
there must
(9.39)
(9.40)
Multiplying both sides by pkm and using the - identity shows that
2G C
pj = irs G rk G rm pkm
(9.41)
or, rearranging and changing the symbols used for the free indices,
G ijC =
1
--2
cca B
ranno
ipq G pr G qs rsj
(9.42)
By writing this out explicitly, we see that the component matrix of the cofactor
tensor equals the cofactor of the component matrix for [ G ] (see Eq. 3.39). Inci
dentally, the cofactor tensor is often called the adjugate
tensor. Unfortunately, some writers also call it the adjoint, but this is a dangerous
misnomer (probably originating from mis-hearing the word adjugate). When
applied to tensors, the term adjoint should be defined to mean the complex
conjugate of the transpose at least thats how it should be defined if you
want your lexicon to agree with definitions used by mathematicians. We use
the term cofactor to side-step the whole ugly issue.
Examination of Eq. (9.42) reveals that the cofactor of the transpose equals
the transpose of the cofactor. Thus, we say that these two operations commute,
and we can write G CT = ( G C ) T = ( G T ) C = G TC .
(9.43)
G C = det ( G ) G T
the
transformed area vector, then substidefine A ( G u ) ( G v ) to be
In continuum mechanics, this equation is known as Nansons formula.
(9.44)
81
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
W ji = W ij ,
(9.45)
from which it follows that W 11 = W 22 = W 33 , W 12 = W 21 , W 23 = W 32 , and
W 31 = W 13 . As will be discussed later, it is common practice to associate a
vector w with a skew-symmetric tensor, defined such that w 1 = W 23 ,
, and w = W . In terms of this vector, the most general form for
w2 = W
31
3
12
the 3 3 matrix of a skew-symmetric tensor is
w3 w2
0
[W ] =
w3
w2 w1
w1
(9.46)
w 3 w 1 w 3 w 2 w 32
(9.47)
Referring to Eq. (7.32), we recognize that this matrix is simply the matrix
associated with the dyad w w . Therefore
If w is the axial vector associated with a skew-symmetric tensor W ,
then W C = w w
(9.48)
In many computational analysis codes, skew-symmetric tensors are not saved
as 3 3 matrices (doing so would be a waste of valuable memory). Instead,
when a subroutine requests a skew-symmetric tensor from the host code, it
will instead be given the three components of the axial vector w . If that rou
tine needs to compute the cofactor of the full skew-symmetric tensor
W , then
G 1 = ---------------(9.49)
det ( G )
Cramers rule is very inefficient for computing the inverse of a large matrix,
but it is perfectly adequate for the 3 3 matrices associated with second-order
tensors in 3D space.
82
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
then (as readily verified by back substitution)
C 1 u v C 1
B 1 = C 1 ------------------------------------
1 + v C 1 u
(9.50)
(9.51)
Derivative of a determinant
Let J denote the determinant of a tensor F :
J detF
(9.52)
Recall that the determinant of a tensor can be obtained by simply taking the
determinant of the matrix of components associated with that tensor:
F 11 F 12 F 13
J det F 21 F 22 F 23
(9.53)
F 31 F 32 F 33
From this we see that J = J ( F 11, , F 33 ) . In other words, J is a function of
the nine components of F . Now we seek to know the partial derivative of J
with respect to any one of the components F ij holding the other components
fixed. Recall from simple matrix analysis that the determinant in Eq. (9.53)
can be computed using cofactors:
3
J =
F ij F ijC
(9.54)
j=1
Now recall that the cofactor F ijC is the signed determinate of the submatrix
obtained by striking out the i th row and j th column of [ F ] . Consequently, F ijC
does not depend on the value of F ij , and the derivative of Eq. (9.54) gives simply
J
----------- = F ijC
F ij
cca B
ranno
(9.55)
dJ
In direct notation, this result is written as, -------- = F C or, if F is invertible,
dF
dJ
-------- = J F T
(9.56)
dF
83
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Applying the definition of a dyad, this can be written
(10.1)
Q( x) = Q x
(10.2)
where
Q = I nn
(10.3)
P( x) = P x ,
(10.4)
P = I Q = nn
(10.5)
where
P = e1 e1 + e2 e2
1 0 0
[ P] = 0 1 0
0 0 0
w.r.t. e 3 =n
(10.6)
84
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Q = e3 e3
Rebe
0 0 0
[Q] = 0 0 0
0 0 1
w.r.t. e 3 =n
(10.7)
merely the sum of dyads of an orthonormal basis for the range space. In general, if a range space is known, all one has to do is obtain an orthonormal
basis for the range space and sum the diagonal basis dyads to obtain the projector. This concept is used extensively in the main part of this report. Keep in
mind that Q is also a projector. Its associated range space in this example is
be written
QQ = Q
PP = P
PQ = QP = 0
P+Q = I
P
Q
x x = x PT Q x
For example, the projection tensor associated with Eq. (6.50) is
ab
Q = I -------------
ab
or, using the no-symbol dyadic notation,
cca B
ranno
(10.8)
(10.9)
85
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
ab
Q = I ----------- ab
The complementary projector is
(10.10)
ab
P = -----------(10.11)
ab
The general properties listed in Eq. (10.8) can be readily verified to hold for
this particular example.
For example, we see that the projector in Eq. (10.11) is self adjoint if and only
if a b = b a . Recalling Eq. (7.25), this is possible only if b = a , in which
the projection
(6.11)
in which
case
reduces to the orthogonal projection of Eq.
the unit vector n is just a a .
A linear self-adjoint projector transforms its argument to the nearest point
on the range space. This means that the null space is orthogonal to the range
space. For ordinary projections in 3D-space, this means that the vector is projected orthogonally, not obliquely. When Eq. (10.12) holds, the last property
listed in Eq. (10.8) becomes
x P xQ = 0
(10.13)
Suppose that two distinct vectors x and y are decomposed via the projection
theorem as
x = x P + xQ
y = y P + yQ
(10.14)
(10.15)
x y = x P y P + xQ y P + x P yQ + xQ yQ
If the projection is self-adjoint, then any vector in Q-space must be perpendicular to any vector in P-space. Thus, the middle two inner products in the
above equation are both zero, and we obtain
x y = x P y P + xQ yQ
Equivalently,
(10.16)
86
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
xy = xPy+xQy
(10.17)
Q = I P . Suppose that the range space associated with one of these projec
has
tors
a dimension greater than one. Then that projector may always be further decomposed into more projectors, all of which may be made
complementary to each other. For second-order tensors referenced to ordinary
3D space, there can be up to three projectors P 1 , P 2 , and P 3 . For second
P i P j = 0 if i j
k=1
Pk = I
(10.18)
If all of the projectors are self-adjoint (i.e., if they are all symmetric) then
dimP i may alternatively be computed by
dimP i = P i :P i
(10.20)
where the double dot operation is defined such that, for any tensors A and B ,
x y = x P k y
k=1
(10.21)
plane. This projection can be decomposed further into more primitive projectors; namely e 1 e 1 and e 2 e 2 . Hence, if desired, we could define a set of three
generalized mutually
orthogonal
complementary projectors as
P1 = e1 e1
1 0 0
[ P] = 0 0 0
0 0 0
cca B
ranno
(10.22)
87
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
P2 = e2 e2
000
[ P2 ] = 0 1 0
000
(10.23)
P3 = e3 e3
000
[ P3 ] = 0 0 0
001
(10.24)
= x e1 e1 y + x e2 e2 y + x e3 e3 y
= x1 y1 + x2 y2 + x3 y3 ,
(10.25)
88
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Since the transformation is linear, there must exist a second-order tensor Q
such that
f ( x) = Q x ,
so Eq. (15.36) becomes
(11.2)
(Q x) (Q x) = x x
(11.3)
or
x ( QT Q I ) x = 0
(11.4)
This must hold for all x and (because the tensor in parentheses is symmetric),
Thus, the transpose of the tensor Q must equal its inverse. A tensor is said to
Equivalently,
Q is orthogonal
QT Q = Q QT = I
cca B
ranno
(11.6)
(11.7)
89
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
( detQ ) 2 = 1
(11.8)
cos sin 0
sin cos 0 with respect to basis where e 3 = a .
0
0 1
(11.9)
0 0 0
001
0 0 0
with respect to basis where e 3 = a .
(11.10)
100
0 1 0
000
[ I a a ] = 0 1 0 and [ a a ] = 0 0 0 , and [ A ] = 1 0 0
000
0 0 0
001
(11.11)
with respect to basis where e 3 = a .
(11.12)
Many people wrongly claim that an improper orthogonal tensor corresponds physically to a reflection, but it generally represents a reflection in
combination with a rotation. In 3D, there are two types of simple reflections:
(1) a reflection about the origin, which merely reverses the direction of all vectors without affecting their magnitudes, or (2) a mirror-like operation that
will transform any vector x to become its mirror image across some plane
defined by a unit normal n. The first type of reflection fully inverts space
90
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
identity tensor, Q = I . For the second type of reflection, the part of the vec
become
n n x . However, the projection of the vector onto the plane of the mirror,
x n n x remains unchanged. Thus, the reflection is given by the sum of this
part plus the reversed part to give
unchanged
f ( x) = x 2(nn x)
The corresponding reflection tensor is
(11.13)
Q = I 2n n
(11.14)
or
1 0 0
[ Q ] = 0 1 0 with respect to basis where e 3 = n .
0 0 1
cca B
ranno
(11.15)
1 0 0 1 0 0
[ Q ] = 0 1 0 0 1 0
0 0 1 0 0 1
(11.16)
with respect to basis where e 3 = n .
Note that the last matrix in this equation is identical to the matrix in
Eq. (11.19) with the rotation angle set to 180 . This is just one special case
of a more general statement. Namely, n 3D, if Q is an improper orthogonal
ab = Ab
(11.18)
The i th component of A b is
( A b ) i = A ij b j
(11.19)
For Eq. (11.18) to hold, the right hand sides of Eqs. (11.17) and (11.19) must be
equal:
91
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A ij b j = ijk a j b k
(11.20)
This relationship must hold for all b . Before we can eliminate b , however, we
subscripts so that the indices
(11.21)
Now that both b s have the same index, we may assert that this holds for all
b j and therefore
A ij = ikj a k = ijk a k ,
(11.22)
( a b )i
b k
( ijk a j b k )
= ----------------------------- = ijk a j --------A in = ----------------------(11.23)
b n
b n
b n
In the last step, we have used the fact that neither ijk nor a j depends on b .
To simplify this result, we apply Eq. (9.10) to write b k b n = kn so that
A in = ijk a j kn
(11.24)
(11.25)
(11.26)
In matrix form, the components of A are
(11.27)
92
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
0 a3 a2
[ A] =
a3
0 a1
a2 a1
(11.28)
Whenever a relationship of the form in Eq. (11.27) hold, we say that a is the
property that
AT = A
(11.30)
ba = b A
(11.31)
rigid body. If the rigid body rotates about the origin with an angular velocity vector , then the
P
v
rigid body
rota
t
axision
v = r
(11.32)
Equation (11.18) implies that this relationship
may be written alternatively in terms of an angular velocity tensor . Namely,
v = r,
(11.33)
where
0 3 2
[] =
0 1
2 1
cca B
ranno
(11.34)
Rotation can be defined in terms of an axis and an angle of rotation, but multiple rotations about different axis do not commute (i.e., if you rotate about one
axis and then about the other, the result will not be the same if you switch the
order of rotation). Rotation can nevertheless be shown to be a linear transfor-
93
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
mation. In other words, given a collection of vectors, you can rotate them all
individually and then take a linear combination of the results or you can first
take the same linear combination of starting vectors and then rotate the result
either way, the final result will be the same. Hence, there exists a tensor R
such that rotating an initial vector r o to obtain a new vector r can be written
(11.35)
r = R ro
If the rotation varies in time, we can take time rates of this equation to obtain
the velocity, v = r . Specifically, noting that the initial vector r o is (obviously)
o
Noting from Eq. (11.35) that r o = R 1 r , Eq. (11.36) may be written
v = r , where R R 1
(11.37)
It can be shown that a rigid rotation tensor must be orthogonal* and hence
(11.38)
R 1 = R T
It can further be shown that this property in turn implies that must be
skew symmetric. The associated axial vector is the angular velocity vector .
Any tensor with this property is said to be skew-symmetric. Here, the superscript T denotes the transpose.
A ij = A ji
(11.40)
1
sym A --- ( A + A T )
(11.41)
2
the skew-symmetric (or antisymmetric) part of A is defined
1
(11.42)
skw A --- ( A A T )
2
* For a detailed discussion of orthogonal tensors, including the definition of Euler angles, and how to
generate a rotation from an axis and angle, see Ref. [5].
94
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
With these definitions, we see that any tensor A can be decomposed addi
tively into symmetric and skew-symmetric parts:
(11.43)
A = sym A + skw A
Instead of sym and skw, many authors use superscripts s and a so that
Eq. (11.43) would be written somewhat more compactly as
s
(11.44)
A = A +A
Note that the act of taking the symmetric part of a tensor may be regarded as
a tensor function:
1
P ( A ) = --- ( A + A T )
2
(11.45)
symmetric part of a tensor that is already symmetric is just the tensor itself.
We can also define a function for taking the skew-symmetric part of a tensor:
1
Q ( A ) = --- ( A A T )
2
(11.46)
The functions P and Q have the properties listed in Eqs. (6.51) except that,
this time, the argument of functions is a tensor rather than a vector. Hence,
the fact that any tensor can be split into symmetric and skew-symmetric parts
is merely a generalized application of the projection theorem!
(11.47)
where : denotes the tensor inner product, soon to be defined in Eq. (12.1).
Noting that the dyad x x is a symmetric tensor, we can use the (yet to be pre
sented) result of Eq. (12.15)
to note that B : x x = ( sym B ): x x . Therefore, a
A sym B
cca B
ranno
(11.49)
95
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
The tensor A (and therefore B ) is positive definite if and only if its associated
component matrix is positive definite. This will happen if and only if all three
characteristic invariants of A are positive. WARNING: It is possible to con
struct a tensor B that has all positive invariants, but the tensor is not positive
out that computing all of these principal minors is not necessary in order to
test for positive definiteness. All you have to do is test one principal minor of
each size, making sure that, as you progress in size that the smaller principal
minors are nested inside of each larger minor. The concept of nesting makes
more sense if we explain it starting with the largest principal minor, which is
the determinant of the 3 3 matrix itself. For the next principal minor, select
any 2 2 submatrix whose diagonal components are also diagonal components of the larger matrix, and verify that the determinant of this 2 2 submatrix is positive. Then test the determinant of any 1 1 submatrix on the
diagonal of the 2 2 matrix (i.e., check any diagonal component of that
matrix). The following sequence of decreasing sized submatrices are all principal submatrices, but they are not nested:
A 11 A 12 A 13
Not nested:
A 21 A 22 A 23
A 31 A 32 A 33
A 11 A 12
A 21 A 22
, [ A 33 ]
(11.50)
det
A 11 A 12
A 21 A 22
(11.51)
>0
(11.52)
96
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
A 11 A 12 A 13
det A 21 A 22 A 23 > 0
(11.53)
A 31 A 32 A 33
Recalling that you may use any set of nested principal minors, it follows that a
positive definite tensor will always have positive diagonal components, so this
is a good thing to visually check before even looking at the larger determinants. There is nothing, that disallows the off-diagonals from being negative,
so long as the principal minors always evaluate to a positive number.
from which it follows that its invariants (or principal subminors) will all be
simply non-negative (i.e., positive or zero).
A tensor C is negative semi-definite if C is positive semi-definite.
denoted variously as A
or A d or A or dev A , and it is defined
dev
1
A A A --- ( tr A )I
(11.55)
3
Here, tr A is a scalar called the trace of A defined
tr A A kk = A 11 + A 22 + A 33
(11.56)
iso
The isotropic part of A is denoted A or iso A , and is defined
iso 1
A --- ( tr A )I
(11.57)
3
cca B
ranno
iso
A = A +A
(11.58)
This is also an application of the projection theorem! This equation is analogous to Eq. (__).
97
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A : B A ij B ij
(12.1)
This is a summation of every component of A multiplied by the corresponding
The double dot product is defined such that it must operate between tensors of at least second-order. There is no need for parentheses in Eq. (12.2)
because b :r would be meaningless the double dot is understood to reach
out until it is acting between two tensors (in this case, a b and r s ). Thus, for
operation would
be
a i b m c n T mn
(12.4)
98
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
Some writers [e.g., Ref. 17*] prefer always using a single raised dot to denote
all inner-products, regardless of the order. These writers demand that meaning of the single-dot operator must be inferred by the tensorial order of the
arguments. The reader is further expected to infer the tensorial order of the
arguments from the context of the discussion since most writers do not indicate tensor order by the number of under-tildes. These writers tend to define
the multiplication of two tensors written side by side (with no multiplication
symbol between them) to be the tensor composition. For example, when they
write AB between two tensors that have been identified as second-order, then
they mean what we would write as A B . When they write UV between two
tensors that have been identified as fourth-order, they mean what we would
write as U :V . Such notational conventions are undeniably easier to typeset,
and they work
fine whenever one restricts attention to the small set of conven* We call attention to this reference not because it is the only example, but because it a continuum
mechanics textbook that is in common use today and may therefore be familiar to a larger audience.
99
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
U ::V
U ijpq V pqkl
U :V
U ijkp V plmn
UV
U ijkl V mnpq
UV
A pq U pqij
A :U
A ip U pjkl
AU
A ij U klmn
AU
A ip pjq U qklm
AU
(12.7)
(12.8)
(12.9)
100
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
The difficulties with direct notation might seem to suggest that perhaps
indicial notation would be the best choice. In some instances, this is true.
However, even indicial notation has its pitfalls, principally in regard to operator precedence. For example, the notation
f
----------- kk
(12.10)
( tr )
cca B
ranno
(12.11)
Bottom line: always employ notation that seems most likely to accomplish
the desired interpretation by the educated reader.
A A: A
(12.12)
This definition has exactly the same form as the more familiar definition of
the magnitude of a simple vector v:
v vv
(12.13)
Though rarely needed, the magnitude of a fourth-order tensor X is a sca
lar defined
(12.14)
X = X :: X
A vector is zero if and only if its magnitude is zero. Likewise, a tensor is zero if
and only if its magnitude is zero.
A : B = dev A :dev B + iso A :iso B
(12.16)
Decomposition of the tensors into its symmetric plus skew symmetric parts
( A = sym A + skw A and B = sym B + skw B ) represents an orthogonal projec
101
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
the symmetric
product between B any other tensor A will depend only on
metric part without any loss in generality which can save on storage in
numerical computations.
Incidentally, note that the trace operation defined in Eq. (11.56) can be
written as an inner product inner product with the identity tensor:
(12.17)
tr A I : A
Also note that I :I = trI = 3 , so Eq. (12.16) may be alternatively written
1
A : B = A : B + --- ( tr A ) ( tr B )
(12.18)
3
102
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
A( B: X )
.
(12.19)
P ( X ) = X ---------------------A: B
As for the projection in 3-space, this operation represents a linear oblique projection in tensor space. The surface to which X is projected is orthogonal to
B and the oblique projection direction is aligned with A . This projection func
tion appears in the study of plasticity [6] in which a trial stress state is
returned to the yield surface via a projection of this form.
The fourth-order projection transformation can be readily verified to have
the following properties:
P ( X ) = P ( X ) for all scalars .
(12.20)
P( P( X )) = P( X ) .
(12.22)
The first two properties simply indicate that the projection operation is linear.
The last property says that projecting a tensor that has already been projected
merely gives the tensor back unchanged.
Finally, the analog of Eqs. (6.26) and (6.27) is the important identity that
P( X ) = P(Y )
X = Y + A.
if and only if
(12.23)
This identity is used, for example, to prove the validity of radial return algorithms in plasticity theory [6].
103
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Therefore,
e appears.
even though x depends on three coordinates { r, , z } ,
laboratory basis,
it does so happen that the vector components are identical to
the vector coordinates { x 1, x 2, x 3 } .
As mentioned above, the choice of basis is almost always motivated by the
choice of coordinates so that each base vector points in the direction of increasing values of the associated coordinate. However, there is no divine edict that
demands that the base vectors must be coupled in any way to the coordinates.
If, for example, you were studying the mechanics of a ferris wheel, then you
might favor using cylindrical coordinates to identify points in space (with the
z axis parallel to the wheels axis, but the laboratory basis (with, say, the e 2
discussion
104
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
as follows:
v = v i e i
v = v i e i
(13.1)
(13.2)
(13.3)
v i e i = v i e i
The hat components { v 1, v 2, v 3 } are not generally equal to the tilde components { v 1, v 2, v 3 } , but they are related in a very specific way.
Because the bases are orthonormal, we have the usual identities that
e i e j = ij
(13.4)
and
e i e j = ij
We can also compute the mixed dot products between the base vectors:
(13.5)
Q ij = e i e j = e j e i
(13.6)
Q ij = cos ij
(13.7)
where ij is the angle between e i and e j . For this reason, the Q ij are some that relate
v i Q ij = v j ,
(13.8)
which shows how the v j components can be computed if the v i components
are known.
Dotting both sides of Eq. (13.1) by e j and applying Eq. (13.5) gives
v j = v i Q ji ,
(13.9)
105
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
{ v } = [ Q ] T { v } and { v } = [ Q ] { v }
(13.10)
The only way that these equations can both be true is if [ Q ] is orthogonal:
[ Q ] 1 = [ Q ] T
(13.11)
This property holds only for transformations between orthonormal bases. Furthermore, if the handedness of both bases is the same, then
det [ Q ] = +1
(13.12)
If the handedness of one basis is different from the other, then det [ Q ] = 1 .
The transformation relationships between tensors can be derived in a similar manner. Specifically, if T = T ij e i e j = T ij e i e j , then
(13.13)
T ij = Q ip Q jq Tpq
and
T ij = Q pi Q qj T pq
(13.14)
Again, note that the first indices on each Q match the indices on the hat
components while the second indices on each Q match the indices on the
tilde components. Many authors use these tensor transformation rules as a
definition of the term tensor.
The transformation relations for a third order tensor h are
h ijk = Q Q Q h
ip jq kr pqr
(13.15)
and
h ijk = Q pi Q qj Q rk hpqr
(13.16)
The transformation relations for tensors of higher order are constructed similarly.
As a final remark, we recall that
(13.17)
Q ij = e i e j
We can always construct a basis-dependent tensor Q as
(13.18)
Q = e k e k
e i Q e j = e i e k e k e j = ik kj = ij
(13.19)
e i Q e j = e i e k e k e j = ik Q kj = Q ij
(13.20)
(13.21)
e i Q e j = e i e k e k e j = Q ik kj = Q ij
106
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Thus, the component matrix for Q is the identity matrix when referenced to a
cca B
ranno
(13.22)
(13.23)
e k = Q e k
and e i = Q ijT e j
This means that the columns of the Q ij matrix are the components of the
tilde basis with respect to the hat basis. Furthermore, the rows of the Q ij
matrix are the components of the hat basis with respect to the tilde basis.
107
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
108
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
bers defined to be two times the index of the number. Then the first number
would be 2, the second 4, and the third 6. He would have an array of three
numbers given by { 2, 4, 6 } . To determine if your neighbor is actually working
with a vector, you would first have to ask him if the indices of his rule had anything to do with a particular set of directions. Since this tutorial limits its
scope to orthonormal bases, we will presume that he told you that the three
directions corresponded to east, north, and up (which are orthogonal). You
would then ask your neighbor if his index doubling rule would still apply if he
were to instead use three different orthogonal reference directions such as
north-east, north-west, and up. If his answer is yes, the doubling rule applies
for any set of orthonormal reference directions, then you would conclude that
he is merely working with an array of three numbers, not a vector. To prove
that your neighbors doubling rule does not correspond to a vector, you would
only have to find one counter example. Let { e 1, e 2, e 3 } denote the directions
the
directions {north-east,
{east, north, up}. Now let { E 1, E 2, E 3 } denote
north-west, up}. In other words, let
e1 + e2
-,
E 1 = ---------------
2
e1 + e2
-,
E 2 = --------------------
2
E3 = e3
(14.1)
(14.2)
4e 1 + 6e 2
+ 6 E3
---------------------------
2
(14.3)
or
Thus, the doubling rule applied in the second system does not result in the
same vector as would be obtained by applying that rule in the first system
(namely, 2e 1 + 4e 2 + 6e 3 ), and you must therefore conclude that your neigh
cca B
ranno
(14.4)
109
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
In other words, even though the components will generally vary upon a change
of basis, the sum of components times base vectors does not change. Thats
why the term invariance is used. It says that a vector (the sum of components
times basis) is invariant under a change of basis. If the basis changes, then
the components must change in a particular way so that Eq. (14.4) remains
true. For second-order tensors, a similar statement applies. Namely, the set of
nine basis dyads ( e 1 e 1, e 1 e 2, e 1 e 3, e 2 e 1, , e 3 e 3 ) forms a basis for second order
numbers
( a ,a , a , a , , a ) corresponds to a sectensors. A set of nine
11 12 13 21
33
ond order tensor if and only if the rule that assigns these numbers changes
upon a change of basis in such a manner that
a ij e i e j = A ij E i E j
(14.5)
where
{ E 1, E 2, E 3 }
is
any
other
(orthonormal)
basis
and
( A 11, A 12, A 13, A 21, , A 33 ) are the nine new numbers resulting from applying
the rule in the new basis. If the above equation does not hold true, then the
rule, and any numbers it spits out, do not correspond to a tensor.
Thankfully, the triplets and ordered 3 3 arrays that pop up in physics
typically do satisfy the vector and tensor invariance requirements of
Eqs. (14.4) and (14.5). Thats no surprise, since any physical phenomenon
should be unaffected by the coordinate frame that we adopt to view it from.
The components do not mean much by themselves, but the sum of components
times bases (being invariant) does have strong physical significance. It is the
sum that means something as a single entity. Even though describing a velocity requires supplying three numbers, youll never hear a person say the
velocity are . Being a vector, velocity is (and should be) regarded as a single
entity. Physical phenomena are routinely described conceptually using direct
notation and practically (i.e., for particular problems or in computer codes)
using components.
Whenever a rule is proved to satisfy invariance, we always endow it with
a direct (Gibbs) notation. Whenever a new operation is defined exclusively in
terms of invariant operations, then the new operation is immediately known
to be itself invariant.
110
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(14.7)
(14.8)
To demonstrate that the rule satisfies tensor invariance, we must demonstrate that Eqs. (14.7) and (14.8) give the same answer for the scalar s . Recalling Eq. (13.9) we know there exists an orthogonal matrix [Q] such that
v j = v i Q ji
(14.9)
Substituting this into Eq. (14.7), being extra careful to introduce two distinct
dummy summation indices, gives
s = v m Q im v k Q ik
(14.10)
(14.11)
which is identical to Eq. (14.8). Thus, this rule for constructing the scalar s
satisfies tensor invariance. Now that we know that this rule is invariant, we
are permitted to give it a direct notation symbol. Of course, the symbol for this
operation is already defined it is simply the dot product:
s = vv
cca B
ranno
(14.12)
{ e 1, e 2, e 3 } , lets consider
the following scalar-valued rule:
s = v1 + v2 + v3 ,
(14.13)
111
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
In other words, the scalar is constructed by simply adding up the vectors components. To show that this rule is not invariant, all we need is a counterexample. Consider a vector
v = 3e 1 + 4e 2
(14.14)
1
1
(14.15)
e 1 = --- ( 3e 1 + 4e 2 ) , e 2 = --- ( 4e 1 + 3e 2 ) , and e 3 = e 3
5
5
v = 5e 1
(14.16)
Now, if we apply the rule of Eq. (14.13) to Eq. (14.14) we obtain s = 7 . However, if we apply the rule to Eq. (14.16), we get s = 5 . The final results do not
agree! Consequently, the component summation rule is not a valid tensor rule.
Note, by contrast, that Eqs. (14.7) and (14.8) do give the same result for this
sample vector.
( x 2 + y 2 ) dV
(14.17)
Here, { x, y, z } are the Cartesian coordinates of a point in the body. Thus, the
quantity ( x 2 + y 2 ) is the square of the distance of a point from the z axis.
Intuitively, the farther points are from the z axis, the harder it is to rotate
those points body about the z axis. Hence, zz is a good measure of the resistance of the body to rotation about the z axis. The moments of inertia about
the other two axes are defined similarly:
xx =
( y 2 + z 2 ) dV
(14.18)
yy =
( z 2 + x 2 ) dV
(14.19)
112
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
xy =
( xy ) dV ,
Rebe
yz =
( yz ) dV ,
zx =
and
( zx ) dV
(14.20)
Though less intuitive, these quantities measure how much wobble you would
feel when rotating the body about an axis.
It would seem natural to arrange these inertias into a 3 3 matrix as
xx xy xz
[ ] = yx yy yz
(14.21)
zx zy zz
Here, we have defined yx = xy , etc. to construct a symmetric matrix. It is
perfectly legitimate for us to arrange the inertias into a matrix form. The
question arises, however:
QUESTION: Does the [] matrix correspond to a tensor?
ANSWER: No, it does not!
Whenever you wish to prove that something is false, all you need is a counterexample. Consider a body that is a tiny ball of mass m located on the x axis
at a distance L from the origin. Treating the ball as a point mass so that
x L , y z 0 , the inertia values are
yy = zz = mL 2 and all other ij 0
(14.22)
Thus,
[] =
mL 2
000
010
001
(14.24)
zz = mL 2
(14.25)
1
xy = --- mL 2
2
(14.26)
(14.23)
y
m
cca B
ranno
x
L
Thus,
113
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
[ ] = mL 2
12 12
12 12
0
0
1
(14.27)
Now the question becomes: would we have obtained Eq. (3.46) if we had
merely performed a coordinate transformation of Eq. (14.23)? Lets show that
the answer is no. The direction cosine matrix for the transformation from
the { xyz } system to the { x y z } system is
1
------2
1
[ Q ] = -----2
1
------2
1
------2
(14.28)
0 0 1
If we assume (incorrectly) that the [ ] matrix corresponds to a tensor, then
the following relationship should be true
[]
w.r.t.
{ x y z }
= [ Q ] T [ ] w.r.t. [ Q ]
(14.29)
{ xyz }
------- ------- 0
------- ------- 0
000
12 12 0
2
2 2
2
This equation is false!
1 1
1 1
0
1
0
1 2 1 2 0 = ------ ------- 0
------- ------- 0
2 2
2 2
001
0
0 1
0 0 1
0 0 1
(14.30)
Multiplying this out, however, shows that equality does not hold. Consequently, this counterexample proves that the matrix of inertias defined in
Eq. (14.21) is not the matrix associated with a tensor!
yx yy yz
(14.31)
zx zy zz
This new definition corresponds to a tensor if and only if, for any two Cartesian coordinate systems,
* Keep in mind: we are trying to prove that it doesnt.
114
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
[]
w.r.t.
{ x y z }
Rebe
[ Q ]T [ ]
w.r.t. [ Q ]
{ xyz }
(14.32)
Lets first check whether placing negatives on the diagonals fixes our previous
counterexample. With our new definition for the inertia tensor, we now ask
whether negatives on the diagonals of the inertia tensors in Eq. (14.30) will
make that equation true. In other words, is the following matrix equation
true?
1
------- ------- 0
------- ------- 0
0 0 0
1 2 1 2 0
2
2 2
2
=
This equation is true!
1
1
1
1
1 2 1 2 0
------- ------- 0 0 1 0 ------- ------- 0
2
2
2
2
0 0 1
0
0 1
0 0 1
0 0 1
(14.33)
Multiplying this out shows that equality holds. Passing this test is necessary,
but not sufficient. he fact that one example worked out okay does not prove
that our new definition really does correspond to a tensor. We dont know for
sure (yet) whether or not some other counterexample might prove that our
new definition isnt a tensor either. It is always harder to prove something is
true than to prove that it is false!
It turns out that our new definition does indeed transform properly for all
possible coordinate changes. One of the simplest ways to prove that something
is a proper invariant tensor rule is to find some way to write out the definition
in direct (Gibbs) notation using only operations (such as the dot product and
dyadic multiplication) that have already been proved to be tensor operations.
We recognize that the expression x 2 + y 2 that appears in the definition of
zz can be written as ( x 2 + y 2 + z 2 ) z 2 , or simply x x z 2 , where the posi observation lets us
tion vector is x = xe x + ye y + ze z = x e x + y e y + z e z . This
into
the formula. Note that the offintroduce at least
a bit of direct notation
diagonals of the dyad x x are identical to the off-diagonals of our new inertia
[ x x ( x x ) I ] dV
cca B
ranno
(14.34)
115
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(15.1)
(15.2)
To prove that the sum of squares is a vector invariant, we must prove that
Eq. (15.2) holds true whenever the v k components are related to v j components through a regular basis transformation, namely v k = Q ik v i where Q jk
are the components of any proper orthogonal matrix. Making this substitution
for v k in the expression v k v k gives
v k v k = Q ik v i Q mk v m = v i Q ik Q mk v m = v i v i ,
(15.3)
116
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
T 11 + T 22 + T 33
= T ij Q ki Q kj
Rebe
= T kk
(now use orthogonality)
= T ij ij
= T ii
= T 11 + T 22 + T 33
(15.4)
Primitive invariants
As mentioned earlier, the magnitude of a vector is an invariant. Likewise,
the square of the magnitude is also an invariant. These two invariants are not
independent. One can always be computed from the other.
Primitive invariants of a vector or tensor are any minimal set of invariants
such that all other invariants may be expressed as functions of the primitive
invariants.
A vector has only one primitive invariant its magnitude. Symmetric tensors (referenced to 3D space) have three primitive invariants. This follows
because symmetric tensors have three eigenvalues. Since an invariant may be
computed in any basis with the same result, all invariants of symmetric tensors must be expressible as functions of the tensors eigenvalues { 1, 2, 3 } .
For example, the magnitude of a symmetric tensor is an invariant that may be
written as 12 + 22 + 32 . Nonsymmetric tensors have more than three primitive
invariants. For nonsymmetric tensors, the magnitude is itself an independent
invariant that cannot be expressed as a function of the eigenvalues. To prove
this statement, consider a tensor whose component matrix with respect to
some particular basis is given by
1 W 0
0 2 0
cca B
ranno
(15.5)
0 0 3
The eigenvalues are { 1, 2, 3 } , but the tensor magnitude is
12 + 22 + 32 + W 2 , which depends on the 12 component W . Different values
of W will give different magnitudes. Hence, the magnitude must not be
expressible as a function of the eigenvalues. Tensor magnitude is an independent fourth invariant for nonsymmetric tensors! Thus, dont let anyone tell
you that a tensor is zero if all of its eigenvalues are zero that statement is
true only for symmetric tensors!
117
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Trace invariants
Three popular invariants of a second-order tensor are
I = trT
II = tr ( T 2 )
III = tr ( T 3 )
(15.6)
(15.7)
Characteristic invariants
The characteristic invariants of a second-order tensor T are defined as fol
lows:
I k = the sum of all possible k k principal subminors
(15.8)
T 11 T 12
T 21 T 22
+ det
T 22 T 23
T 32 T 33
+ det
T 11 T 13
T 31 T 33
T 11 T 12 T 13
I 3 = det T 21 T 22 T 23
(15.9)
T 31 T 32 T 33
118
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
These are called the characteristic invariants because they are coefficients in
the characteristic equation that gives the eigenvalues of T . Namely,
3
2
(15.10)
I1 + I2 I3 = 0
This equation is true even if the tensor T is nonsymmetric. For symmetric
(15.11)
Thus, comparing with Eq. (15.7), we note that the characteristic invariants
are related to the trace invariants by
tr ( T ) = I 1
tr ( T 2 ) = I 12 2I 2
tr ( T 3 ) = I 13 3I 1 I 2 + 3I 3
(15.12)
These relationships hold only for 3D tensors. However, the basic definition of
the characteristic invariants and their relationship with the characteristic
equation extends analogously to other dimensions. For 4-dimensional space,
the characteristic equation is 4 I 1 3 + I 2 2 I 3 + I 4 = 0 . For two-dimensional space, the component matrices are only 2 2 and the characteristic the
characteristic equation is
2 I 1 + I 2 = 0 ,
where
I 1 = T 11 + T 22
and
I 2 = det
T 11 T 12
T 21 T 22
cca B
ranno
(15.13)
Applying this formula is the fastest and least error-prone way to write down
the characteristic equation of a 2 2 matrix.
Direct notation definitions of the characteristic invariants. The
invariant definitions that we have given so far demand that the underlying basis
must be orthonormal. Some more advanced texts (e.g., [20]) cite more
abstract, and therefore more broadly applicable, definitions of the characteristic invariants. These definitions are phrased in terms of the triple scalar product, [ ( ), ( ), ( ) ] , defined in Eq. (5.34). Namely, in 3D, I 1 , I 2 , and I 3 are
defined such that, for any vectors, u , v , and w ,
119
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
[ T u, v, w ] + [ u, T v, w ] + [ u, v, T v ] = I 1 [ u, v, w ]
(15.14)
[ u, T v, T w ] + [ T u , v, T w ] + [ T u, T v, w ] = I 2 [ u, v, w ]
(15.15)
(15.16)
[ T u , T v, T w ] = I 3 [ u, v, w ]
Note that the left hand side in the equation for I 1 sums over every possible
way for the tensor T to operate on exactly one of the vectors in the triple sca
lar product. Similarly, I 2 involves sums over every possible way for the tensor
to act on two of the vectors and I 3 involves sums over every way (only one
way) for the tensor to act on all three vectors in the triple scalar product.
Admittedly, these definitions are rather strange looking, and we have committed the proof that the right hand side can, in every instance simplify to one
invariant number times the triple scalar product [ u, v, w ] . Nonetheless,
it is essential to
expressions like these show up occasionally in analyses, and
recognize how beautifully they simplify. Furthermore, these definitions are a
nice place to start when attempting to deduce how to compute the tensor
invariants for irregular (non-orthonormal) bases.
The cofactor in the triple scalar product. We close this discussion with a
final identity, involving the cofactor tensor. In particular, recall the direct
notation definition of the cofactor, given in Eq. (9.38):
for all vectors u and v
GC ( u v ) = ( G u ) ( G v )
[ GC ( u v ) ] w = [ ( G u ) ( G v ) ] w
(15.17)
(15.18)
product [ u, v, G CT w ] . Thus
[ G u, G v, w ] = [ u, v, G CT w ]
(15.19)
This result shows that, when a tensor acts on two arguments in a triple scalar
product, it can be recast as the transpose of the cofactor acting on the previously untransformed third vector. By the cyclic property of the triple scalar
product, we can assert that this statement holds true when the tensor is acting on any two vectors in a triple scalar product.
Applying the above identity to the left-hand-side of Eq. (15.15), and then
applying the definition of trace given in Eq. (15.14) shows that
I 2 = trT c
(15.20)
120
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
enough because the trace of a sum is the sum of the traces. Thus
I 1A + B = tr A + tr B = I 1A + I 1B
Solving Eq. (15.12b) for I 2 gives a formula for the second invariant:
1
I 2 = --- ( I 12 tr ( T 2 ) )
2
1
I 2A + B = --- ( ( tr A + tr B ) 2 tr ( A + B ) 2 )
2
which simplifies to
I 2A + B = I 2A + I 2B + I 1A I 1B tr ( A B )
Without proof, the determinant of a sum can be written
I 3A + B = det ( A + B ) = I 3A + I 3B + Ac : B + A : B c
cca B
ranno
(15.21)
(15.22)
(15.23)
(15.24)
(15.25)
CASE: invariants of the sum of a tensor plus a dyad. Now suppose that
the tensor B is actually a dyad:
B = uv
(15.26)
Then
I 1B = u v
B
I2 = 0
(15.27)
I 3B = 0
(15.29)
(15.28)
121
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
where I is the identity tensor, and 0 is the zero tensor. This theorem is trivial
T4
= I1T 3 I2T 2 + I3T
= ( I 12 I 2 )T 2 + ( I 3 I 1 I 2 )T + I 1 I 3 I
(15.31)
obtain
T 2 I1T + I2 I
T 1 = ------------------------------------I3
(15.32)
( A + u v ) 1 = A 1 ------------------------------------------
1 + v A 1 u
(15.33)
Eigenvalue problems
Consider a general tensor, A . A vector p is called a right eigenvector of
[ A I ] p = 0
(15.35)
122
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
The only way that this equation can hold is if the tensor in brackets is singular. Hence, must satisfy
det [ A I ] = 0
(15.36)
(15.37)
where I i are the invariants of Eq. (15.9). Once an eigenvalue is found, the corresponding eigenvector is determined by enforcing the definition of
Eq. (15.34). For two-dimensional problems, a graphical tool called Mohrs circle, is also very useful for performing eigenvalue analyses.*
Because the equation is cubic, there are up to three distinct eigenvalues,
{ 1, 2, 3 } . Hence there is the potential for having up to three associated
eigenvectors, { p 1, p 2, p 3 } , but there might be fewer (as explained below). Any
contains the proof that the eigenvectors corresponding
linear algebra textbook
to distinct eigenvalues will be linearly independent. In other words,
If i j , then p i and p j are linearly independent.
cca B
ranno
(15.38)
123
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
vector associated with each eigenvalue. For repeat roots, the geometric multiplicity is always less than or equal to the algebraic multiplicity. The
distinction between geometric and algebraic multiplicity is important only for
nonsymmetric tensors because (as any good matrix analysis will prove),
For a symmetric tensor, the geometric multiplicity always
equals the algebraic multiplicity.
Consequently, a symmetric tensor will always have a complete set of linearly
independent eigenvectors that may be used as a basis for 3D space. Such a
basis is sometimes called an eigenbasis.
For repeated eigenvalues, the associated eigenvectors do not have unique
directions. If the geometric multiplicity is two, then there are exactly two associated linearly independent eigenvectors. These two eigenvectors may be used
to define a plane in 3D space. The eigenvectors associated with the eigenvalue
of geometric multiplicity two are unique to the degree that they must lie in
that plane, but their directions (and, of course, magnitudes) are arbitrary. If
desired, they may be scaled such that they are any two orthonormal vectors in
the plane. These statements apply even to nonsymmetric tensors. For symmetric tensors, in light of Eq. (15.39), we note that
A symmetric tensor always has an orthonormal eigenbasis.
(15.40)
(15.41)
Although not guaranteed, it may so happen that a particular non-symmetric tensor of interest also has all of its geometric multiplicities equal to the
algebraic multiplicities. In such a case, that tensor also possesses a spanning
set of eigenvectors { p 1, p 2, p 3 } that may serve as a basis for 3D space. How
ever,
If a non-symmetric tensor happens to have an eigenbasis,
then the eigenbasis will not be orthogonal.
(15.42)
124
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Diagonalizable tensors
Whenever a tensor possesses a spanning set of eigenvectors, { p 1, p 2, p 3 }
(i.e., when a tensor has an eigenbasis), then the tensor is said to be diagonalizable. In matrix analysis, this means that the matrix [ A ] ee containing the
components of A is with respect to the laboratory basis { e 1, e 2, e 3 } is similar
diagonal.
to a diagonal matrix [ ] containing the eigenvalues on the
In other
words, there exists a matrix [ L ] such that
[ A ] ee = [ L ] [ ] [ L ] 1
(15.44)
where
1 0 0
[] =
0 2 0
(15.45)
0 0 3
Comparing Eq. (15.34) with (15.44), we note that the columns of the matrix
[ L ] must contain the components of the three eigenvectors { p 1, p 2, p 3 } with
respect to the orthonormal laboratory basis. If the tensor A is symmetric,
the
where the eigenvectors are here presumed to satisfy the normalization of
Eq. (15.43) so that, indeed, A p i = i p i (no sum on i ). Incidentally, note
that
pi A = i pi
(15.47)
The dual vectors p i , are sometimes called the left eigenvectors by virtue of
cca B
ranno
(15.48)
125
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Eigenprojectors
Consider a diagonalizable tensor A for which two eigenvalues are repeat
A = 1 p1 p 1 + 2 ( p2 p 2 + p3 p 3 ) ,
(15.49)
A = 1 P1 + 2 P2 ,
(15.50)
where
P1 p1 p 1
P2 p2 p 2 + p3 p 3 .
By virtue of Eq. (15.43) we note that
(15.51)
and
P1 P1 = P1
P2 P2 = P2
(15.52)
(15.53)
(15.54)
P1 + P2 = I
Comparing these results with Eq. (10.18) reveals that the eigenprojectors are
complementary projectors and the number M in Eq. (10.18) is equal to the
number of distinct eigenvalues. Because 2 = 3 , the directions of eigenvectors p 2 and p 3 are not unique they only need to reside in the plane defined
If all of the eigenvalues of the tensor instead had been distinct, we could
have performed a similar analysis to obtain
A = 1 P1 + 2 P2 + 3 P3 ,
(15.55)
where
P1 p1 p 1
P2 p2 p 2
P3 p3 p 3
(15.56)
and the properties of Eq. (10.18) would again hold, this time with M=3 .
126
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
P i P j = P i if i= j so each P i is a projector
(15.57)
P i P j = 0 if i j
(15.58)
(15.59)
P1 + P2 + P3 = I
Likewise, if all three eigenvalues had been equal (triple root), then the number of distinct eigenvalues would be M=1 , and the same process would (trivially) give
A = I ,
(15.60)
127
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(16.1)
x1 n1 + x2 n2 + x3 n3 = d
(16.2)
(16.3)
In this case, u i equals the intercept of the plane on the i th axis of the Cartesian coordinate triad.
If the plane does pass through the origin, then the equation of the plane is
just
xn = 0
(16.4)
The equation of a plane passing through a particular point p is therefore
( x p) n = 0
(16.5)
128
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Equation of a line
Consider a straight line that passes through one point x 1 and a second
x that extend
point x 2 . The easiest way to define the set of position vectors
in parametfrom the origin to some point on the line is to write the equations
ric form by simply expressing the fact that the vector from x 1 to any point on
the line must be parallel to the vector from x 1 to x 2 . Letting the proportional
for a line can be written
ity factor be denoted by t, the parametric equation
( x x1 ) = t ( x2 x1 )
(16.6)
Since this is a vector expression, it represents a set of three separate equations. If the parameter t equals zero, then x is located at x 1 . If t=1 , then the
and will result in position vectors outside the line segare permissible
as well,
ment from x 1 to x 2 .
where
y = n
cca B
ranno
(16.7)
x2 x1
- , and L = x 2 x 1
(16.8)
y x 1--2- ( x 1 + x 2 ) , n ----------------L
The parameter varies from L/2 to +L/2 as x varies from x 1 to x 2 .
More generally, one can define a line by any point p on the line and the
be written
orientation n of the line. Then the equation of the line can
(16.9)
y = n ,
where
y = x p,
(16.10)
and n defines the orientation of the line
= yn
(16.11)
This expression for may be substituted back into Eq. (16.9) to give a nonparametric version of the equation of a line:
y = ( y n )n
(16.12)
Physically, this says that the projection of y onto n equals y itself. Even
equation,
a vector
so it really
though Eq. (16.12) is non-parametric, it is still
represents three simultaneous equations. These three equations are not independent. Given arbitrary choices for y 1 and y 2 , it is not generally possible to
129
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
solve Eq. (16.12) for y 3 . Certain solvability conditions must be met in order
for a solution to exist; namely, the above equation may be solved for y 3 if and
only if y 1 n 2 = y 2 n 1 . This solvability condition is expressing the requirement
that y 1 and y 2 must fall on the shadow cast by the line onto the 1-2 plane.
Equation (16.12) is non-parametric, but it is still a set of three distinct
equations. It can be expressed as a single equation by noting that two vectors
a and b are equal if and only if ( a b ) ( a b ) = 0 . Therefore, an equiva
lent
version
of Eq. (16.12) is
y y = ( y n )2
(16.13)
Any of the above boxed equations may be used as an equation of a line in
3D space. All but the last equation are vector expressions, so they actually
represent a set of three equations, each of which involves a linear combination
of the position vector components. The last equation is a single (scalar-valued)
expression for a line, but the price paid for having a single equation is that the
position components appear in quadratic form pretty odd given that it represents the equation for a linear (straight) line!
(16.14)
1 2 3
(16.15)
If, however, the principle axes are not coincident with your basis, then the
equation for the ellipsoid takes on a more general quadratic form. Namely,
xBx = 1 ,
where the tensor B is defined by the inverse of the dyad sum
A = ak ak
(16.16)
(16.17)
130
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Since the a k are (by premise) mutually perpendicular, we know that they
must be eigenvectors
of the tensor A . Consequently, the eigenvalues of the
3
ak ak
B = ------------(16.18)
k4
k=1
a2 = 4 ( e1 + e2 )
a 3 = 3e 3
(16.19)
(16.20)
(16.21)
(16.22)
2 = 4 2
(16.23)
3 = 3
(16.24)
-->
a 2 a 2 = 16 ( e 1 e 1 e 1 e 1 e 1 e 2 + e 2 e 2 )
a 1 a 1 = 9e 3 e 3
-->
000
000
009
110
110
000
-->
(16.25)
16 16 0
16 16 0
0 0 0
(16.26)
(16.27)
Thus,
110
16 16 0
000
1
1
1
[ A 1 ] = -------------4- 1 1 0 + -----------------4- 16 16 0 + ----4- 0 0 0
3
( 2)
(4 2)
000
0 0 0
009
(16.28)
or
17
-----64
1
[ A ] = 15
-----
64
15
-----64
17
-----64
(16.29)
cca B
ranno
1
--9
131
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
15
17 2 17 2 1 2 15
------ x 1 + ------ x 2 + --- x 3 + ------ x 1 x 2 + ------ x 2 x 1 = 1
64
9
64
64
64
(16.30)
Of course, the last two terms, which come from the off-diagonals, may be combined, but we left them separated to emphasize the structure of the solution.
(16.31)
(16.32)
1 = 2 = r
(16.33)
where
(16.34)
Here, e i is the unit vector in the direction of a i and n is the unit vector in the
(16.35)
may be presumed symmetric (if not, then it must be symmetrized in order for
the comments in this section to remain true).
132
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
hyperboloid.
(16.37)
Expanding this out, recognizing the B may be presumed symmetric without
loss, gives
= x B x 2p B x + p B p d
Equating this expression with Eq. (16.37) gives
(16.39)
B = A
(16.40)
2 p B = b
(16.41)
pB pd = c
(16.42)
The last two equations can be solved for p and d . Thus we have the result:
If = x A x + b x + c
then = ( x p ) A ( x p ) + d ,
1 1
1
where p = --- A b and d = --- b A 1 b c
2
4
cca B
ranno
(16.43)
1
1
1
= x + --- A 1 b A x + --- A 1 b + --- b A 1 b c ,
2
4
2
(16.44)
133
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
1
x = --- A 1 ( y b )
2
(16.45)
(16.46)
(16.47)
Suppose that =0 . Then, if this result were for scalars rather than tensors, it
would read
1
0 = --- [ y 2 + b 2 ] 4c ,
a
y = b 2 4ac
or
(16.48)
Thus, we see that the vector y plays a role similar to that played by the dis As a matter of fact, note that Eq. (16.46) can
criminant in scalar quadratics!
be written
x = ( 2 A ) 1 ( b + y ) ,
(16.49)
b b 2 4ac
which is the analog of the scalar quadratic formula, x = ---------------------------------------- . In
2a
the scalar quadratic equation, there were only two values of y that would
make =0 . By contrast, for the tensor case, setting =0 gives
0 = A 1 : [ y y + b b ] 4c
(16.50)
b A 1 b 4c
(16.51)
which describes a quadric surface (i.e., a family of y vectors exists that will
make =0 ).
134
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(17.1)
(17.2)
(17.3)
[ F ] = [ V ] [ r ] , where [ V ] [ Q ] [ ] [ Q ] T and [ r ] [ Q ] [ R ]
(17.4)
or
(17.5)
(17.6)
[ F ] = [ r ] [ U ] , where [ U ] [ R ] T [ ] [ R ] and [ r ] [ Q ] [ R ]
(17.7)
or
Note that [ r ] is itself orthogonal and that [ U ] and [ V ] are symmetric. Thus,
we conclude the Q-R decomposition guarantees the existence of symmetric
matrices [ U ] and [ V ] as well as an orthogonal matrix [ r ] such that
[ F ] = [r][U ] = [V ][r]
cca B
ranno
(17.8)
135
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
F = RU = VR
(17.9)
A tensor is called a stretch if it is symmetric and positive definite. Physically, a
stretch describes a deformation in which there is an orthonormal triad of
material vectors that change length but not orientation. These special vectors
are the principal directions of the stretch and the corresponding eigenvalues
are the ratios of deformed length to undeformed length. A physical interpretation of the polar decomposition is shown in Fig. 17.1. Note that R U repre
sents a stretch U followed by a rotation R .
In the polar decomposition theorem, the descriptors symmetric and positive definite are requirements, not consequences. A stretch must be both
symmetric and positive definite merely showing that a tensor F is symmet
ric is not sufficient to prove that it is a stretch. For example, any rotation of
exactly 180 will produce a symmetric F tensor, but it will not be positive def
inite.
The procedure for finding R , U , and V is as follows:
1
/
2
C = C 1 P 1 + C 2 P 2 + C 3 P 3 and U = C
= C1 P1 + C2 P2 + C3 P3 .
1
Compute R = F U .
Compute V = R U R T .
Alternatively, if only V is desired, it may be computed by V = ( F F T ) 1 / 2 .
* If p k is the eigenvector associated with C k , then the eigenprojector P equals the dyad
k
p k p k , where there is no sum on k.
By computing U in the principal basis of C , we have ensured that it will be symmetric. In general,
an N N positive definite symmetric matrix like C can have an infinite number of square roots,
N
of which only 2 are symmetric and only one is symmetric and positive definite.
136
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Compress the
material by half
along the ellipse
1-axis.
E2
1U
2U
First rotate
the material
counterclockwise
by 60 degrees.
then rotate
the material
counterclockwise
by 60 degrees.
g
1
g
2
E1
1V
2V
Figure 17.1. Visualization of the polar decomposition. This figure shows that the deformation F can be visualized as two step process. The upper R U path first applies the stretch
of the vector labeled U and
U to compress the material by a factor of 1 2 in the direction
1
then
the rotation tensor R rotates counterclockwise by 60 . The same deformation
is
achieved on the bottom path by first rotating by 60 and then compressing by a factor of 1 2
in the direction of the vector labeled 1V . In these figures, we have painted a circle (or
sphere in 3D) on the reference cube to show how it becomes an ellipse (or ellipsoid in 3D).
The vectors kU and kV lie on the major axes of the ellipse.
cca B
ranno
137
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
which is sometimes more computationally convenient.
(17.10)
1
/
2
T
P( X ) = +( X X )
(17.12)
We have inserted the + sign in front of this equation to reiterate that the
square root of a tensor is not unique, but the positive definite square root is
unique. The transformation P is a projection in the sense defined in Eq. (6.36)
because P ( P ( X ) ) = P ( X ) . In other words, if F is already a symmetric posi
follows:
R = F U 1 = F [ + ( X T X ) 1 / 2 ] 1 = Q ( F )
(17.13)
1
/
2
T
Q( X ) = X ( X X )
(17.14)
138
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
, where F 33 > 0
(17.15)
0 F 33
[ R] =
cos sin 0
sin cos 0 ,
0
0 1
[ U ] = [ R ]T [ F ]
and
(17.16)
where
F 11 + F 22
cos = ------------------------------------------------------------------------( F 11 + F 22 ) 2 + ( F 21 F 12 ) 2
F 21 F 12
sin = ------------------------------------------------------------------------( F 11 + F 22 ) 2 + ( F 21 F 12 ) 2
(17.17)
Beware! You must define cos and sin separately in order to uniquely determine the rotation angle. It is certainly true that
tan = ( F 21 F 12 ) ( F 11 + F 22 ) ,
cca B
ranno
(17.18)
but this relation does not uniquely define the rotation angle because there are
always two angles in the range from 0 to 2 that satisfy the above equation.
By contrast there is only one angle in the range from 0 to 2 that satisfies
Eq. (17.17). The following code fragment may be used to compute a polar rotation tensor in 2-D:
c = F(1,1)+F(2,2); s = F(2,1)-F(1,2);
d = Sqrt(c*c+ s*s);
c=c/d;s=s/d;
R(1,1)=c;R(2,1)=-s;R(1,2)=s;R(2,2)=c;
* Its easy to verify that our formulas yield an orthogonal [R] and a symmetric [U]. It is straightforward, but tedious, to also prove that our formula gives a positive definite [U] matrix. This property
is essential in order to confirm the validity of our serendipitous formulas.
139
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
140
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(18.1)
Since this must hold for all orthogonal [ Q ] , it must also hold for any particular choice for an orthogonal [ Q ] . Choosing
0 1 0
[ Q ] = 1 0 0
0 0 1
(18.2)
(18.3)
This condition implies that the tensor must be a scalar times the identity
tensor I . In other words, the most general form for an isotropic tensor refer
enced to 3D space is
0 0
0 0
0 0
(18.4)
where is an arbitrary scalar. Isotropic tensors in 3D space is therefore onedimensional since only one scalar is needed. The identity tensor I is a basis
for this space. Any general tensor B may be projected to its isotropic
part by
the operation
I ( I : B)
iso B = -----------------I :I
Note that I : B = tr B and I :I = trI = 3 . Hence,
1
iso B = --- ( tr B )I
3
cca B
ranno
(18.5)
(18.6)
This is a very familiar result. The idea of finding the isotropic part by projecting to the space of isotropic tensors becomes less obvious when considering
tensors in spaces of different dimensions.
141
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(18.7)
The most general form for an isotropic second-order tensor in two dimensions
therefore contains two arbitrary parameters. Any isotropic tensor in 2D space
may be expressed as a linear combination of the following two primitive base
tensors:
[I] = 1 0
01
and
[] = 0 1
1 0
(18.8)
1
1
= --- I ( I :F ) + --- ( :F )
2
2
I ( F 11 + F 22 ) ( F 12 F 21 )
- + ------------------------------= -------------------------------2
2
(18.9)
In component form,
1 F + F 22 F 12 F 21
isoF = --- 11
2 F F F +F
21
12
11
22
(18.10)
F + F 22 F 12 F 21
1
R = -------------------------------------------------------------------------- 11
( F 11 + F 22 ) 2 + ( F 12 F 21 ) 2 F 21 F 12 F 11 + F 22
for 2D space only!.
(18.11)
142
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
(18.12)
For a proof, see, for example, Ref. [1]. The above expression has three arbitrary parameters, so the space of isotropic fourth-order tensors is threedimensional,* with base tensors given by the parentheses above. The basis
tensors in the parentheses of Eq. (18.12) are only one choice for the basis. In
material modelling, a more convenient basis for the space of isotropic fourthorder tensors is
iso = 1
--- ij kl
P ijkl
3
1
symdev = 1
--- ( ik jl + il jk ) --- ij kl
P ijkl
2
3
skew = 1
--- ( ik jl il jk )
P ijkl
2
(18.13a)
(18.13b)
(18.13c)
Note that these are all constructed from linear combinations of the primitive
bases of Eq. (18.12). Even though the component formulas for this basis are
considerably more complicated, the properties of this basis are irresistible.
Specifically, the basis of Eq. (18.13) consists of complementary projectors! By
this we mean
iso P iso
iso
P ijkl
klmn = P ijmn ,
iso P symdev = 0 ,
P ijkl
klmn
iso P skew = 0
P ijkl
klmn
(18.14)
iso
symdev P iso
P ijkl
klmn = 0 ,
symdev P skew = 0
P ijkl
klmn
(18.15)
skew P iso
P ijkl
klmn = 0 ,
skew P symdev = 0 ,
P ijkl
klmn
(18.16)
Recall that second-order tensors dont really take on any meaning until
they act on a vector. Likewise, the meaning of a fourth-order tensor should be
inferred by operating on a second order tensor. Specifically, for any tensor B ,
we note
1
iso B
-B
P ijkl
kl = -3 kk ij
(18.17)
1
symdev = 1
--- ( B ij + B ji ) --- B kk ij
P ijkl
2
3
(18.18)
1
skew B
- ( B B ji )
P ijkl
kl = -2 ij
(18.19)
* The space is two dimensional if one imposes a minor symmetry restriction that c ijrs =c jirs =c ijsr .
Thats why isotropic elastic materials have only two independent stiffness moduli.
143
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
iso returns the isotropic part of B , P symdev returns the symmetricThus, P ijkl
ijkl
skew returns the skew-symmetric part of skew-symdeviatoric part of B , and P ijkl
metric part of B .
Transverse isotropy
A material property is said to be transversely isotropic about a given unit
vector a if the material can be rotated about this vector without changing the
property. If the property is a scalar invariant, then by definition it
material
cannot change when the material is rotated. The concept of transverse isotropy becomes increasingly complicated for vector and higher order tensor
properties.
In what follows, we often consider a material orthogonal basis
{ p 1, p 2, p 3 } such that p 3 = a . This basis will not generally coincide with the
basis { e , e , e } used in general calculations. Components with
laboratory
1 2 basis
3
respect to the material
are called material components. Components
with respect to the laboratory basis are called laboratory components. Our
goal in what follows is to show the simplified material components of transversely isotropic vectors and tensors and then to show how those tensors may
be expressed directly in the laboratory components without having to perform
a coordinate transformation.
A vector u is transversely isotropic if and only if its material components
u
3
(18.20)
In other words, the only vector that remains unchanged when rotated about
the material symmetry direction a is one that is itself proportional to a .
a3
(18.21)
(18.22)
144
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
At As 0
[ A] =
As At
Aa
(18.23)
Here, the subscripts t and a indicate transverse and axial (relative to the
symmetry axis), and the subscript s indicates shear. If the tensor is symmetric, then it cannot have shear terms and its most general form is
At 0
[ A] =
0 At 0
0 0 Aa
(18.24)
In other words, the tensor must be diagonal in the material basis, with the 11
and 22 transverse components equal. Thus, transverse isotropy means the
tensor must be isotropic in the transverse plane. In dyadic notation, the above
equation may be written
(18.25)
A = At ( p1 p1 + p2 p2 ) + Aa p3 p3
Recall that Eq. (9.16) is true in any orthogonal system; therefore it holds for
the material coordinate system and Eq. (18.25) can therefore be written
(18.26)
A = At ( I p3 p3 ) + Aa p3 p3
or, since p 3 = a
A = At ( I a a ) + Aa a a
(18.27)
Rearranging a bit gives the most general direct notation form for a transversely isotropic second-order tensor.
(18.28)
001
a3 a1 a3 a2 a3 a3
cca B
ranno
(18.29)
Furthermore, the Eq. (18.28) is more useful when the transverse axis is
changing dynamically because Eq. (18.28) can be differentiated with respect
to time.
145
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
A field R must exist. (For this report, the field is the set of reals; for an excellent
definition of the term field, see Ref. [28].)
A2.
A5. There must be a well defined process for determining whether two members of
V are equal.
A3.
A6.
The multiplication and addition rules must satisfy the following rules:
v + w = w + v and v = v
u + (v + w) = (u + v) + w
There must exist a zero vector 0 V such that v + 0 = v .
146
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
A7. ( a, b ) = ( b, a )
A8. ( a, a ) > 0 if a 0 and ( a, a ) = 0 only if a = 0 .
An inner product space is just a vector space that has an inner product.
Ordinary engineering vectors form an inner product space. For engineering vectors, the inner product is computed by
( a, b ) a b = a 1 b 1 + a 2 b 2 + a 3 b 3 = a k b k
For second-order tensors, the inner product is
(19.1)
( A, B ) A : B = A ij B ij
(19.2)
The choice of inner product is not unique. For many engineering applications,
it is often convenient to define an alternative energy inner product that uses a
symmetric positive definite weighting tensor W :
( a, b ) a W b
(19.3)
Note that it is essential that W be symmetric in order to satisfy A7 and posi
tive definite to satisfy A8. Physically,
the word energy is an apt descriptor for
the second-order analog of Eq. (19.3). For continuum elasticity problems, a
material having a strain and a stress has an internal energy given by
u = 1--2- : E : ,
(19.4)
where E is the fourth-order elastic stiffness tensor. Thus, 1--2- E plays a weigh
ing rolemuch like the tensor W in Eq. (19.3).
fg =
f ( x ) g ( x ) dx
cca B
ranno
(19.5)
* The term binary is just an obnoxiously fancy way of saying that the function has two arguments.
147
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Note the similar structures: In (19.1), there is a summation over the dummy
subscript, k. In (19.5), there is an integration over the dummy argument, x.
Once something is discovered to be an inner product vector space, a whole
wealth of already-proved theorems becomes available. For example, the angle
between two vectors is defined by
ab
(19.6)
cos = ---------------------------------
aa bb
We could likewise use Eq. (19.5) to define the angle between two continuous
functions. Recall that the dyad between two ordinary vectors a and b is
defined to be a new object such that ( a b ) v = a ( b v ) for all vectors v .
two
real
continuous
(19.7)
zi =
W ij v j
(19.8)
j=1
z( x) =
W ( x, y )v ( y ) d y
(19.9)
148
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
0 0 1
(19.11)
Then
A :A = A 11 A 11 + A 12 A 12 + + A 33 A 33 = 1 + 9 + 16 + 4 + 1 = 31 > 0
but
(19.12)
A A = A 11 A 11 + A 12 A 21 + + A 33 A 33 = 1 12 12 + 4 + 1 = 18 < 0 , (19.13)
which proves A B is not acceptable as an inner product.
Both A :A and A A are scalar invariants of the tensor A . Note that
0 0 0
cca B
ranno
(19.14)
149
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
Vector subspaces
Consider a set S that is a proper subset of a vector set V . The set S is
called a subspace if any linear combination of members of S is itself in S . A
subspace S is itself a vector space because all of the vector space axioms that
were proved for the larger vector set V continue to hold for set S . Proving that
something is a subspace really only requires proving axioms A3 and A4 with
axiom A2 replaced by the more restrictive discerning definition for membership in S .
Again, a subspace is itself a vector space. It is generally of lower dimension
than V . In ordinary 3D space, a subspace is like a plane any linear combination of vectors in a plane is itself in the plane. By contrast, consider the set
of all unit vectors is not a subspace because a linear combination of two unit
vectors is not itself a unit vector. Recalling that second-order tensors are
themselves nine-dimensional vectors, we encounter tensor subspaces as well.
The set of all symmetric second-order tensors is a tensor subspace because
any linear combination of symmetric tensors is itself a symmetric tensor. The
set of all orthogonal tensors is not a tensor subspace because a linear combination of orthogonal tensors is not itself orthogonal. The set of all transversely
isotropic tensors is not a subspace, but the set (which we will denote TIa) consisting of all tensors that are transversely isotropic about a particular direction a does form a subspace. Specifically, referring to Eq. (18.28) we note that
the tensors, I and a a form a basis for TIa. This basis, is not however, orthog
Q = aa
(19.15)
These tensors form an orthogonal basis for TIa. The basis is orthogonal
because, recalling that a is a unit vector, P :Q = 0 . Furthermore, note that
P :P = 2 and Q :Q = 1
(19.16)
150
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
Hence, this basis is orthogonal, but not normalized. We could naturally divide
P by 2 to normalize the basis, but doing so would make the base tensor no
e i e j . In general, any N th -order tensor may be expressed as a sum of compo (having a total of N indices) times N dyadically multiplied base vectors.
nents
* This section is rather mathematical and it may be skipped without loss in continuity.
151
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
The contraction operator C IJ is defined to dot the I th base vector with the
base vector. Here, I and J are not tensor indices they merely indicate
the positional location of a dyadically multiplied base vectors. The contraction
operation is well-defined only when operating on second- and higher-order
tensors. For example
J th
C 12 ( a b ) = C 12 ( a i b j e i e j )
C 12 ( T ) = C 12 ( T ij e i e j ) =
2
2
C 1 = C 1 ( ijk e i e j e k )
C 13 = C 13 ( ijk e i e j e k )
= a i b j e i e j = a i b j ij = a i b i = a b
T ij e i e j = T ij ij = T ii = trT
= ijk e i e j e k = iik e k
= ijk ( e i e k )e j = iji e j
(19.17)
The contraction operator reduces the order of the operand by two. If, for
example, the operand is a second-order tensor, then the result is a zero-order
tensor, which is a scalar. As seen above, the contraction of a third-order tensor
gives a first-order tensor, which is a vector. The contraction of a fourth-order
tensor is a second-order tensor.
JL dots the I th base vector with the J th vector
The generalized operator C IK
and the K th base vector with the L th vector. This operation reduces the order
by four. To be well defined, all four of the indices, I, J , K , and L , must have
distinct values. The operator is well-defined for tensors of fourth-and higher
order. For example,
42 ( H ) = C 42 ( H
(19.18)
C 13
13
ijkl e i e j e k e l ) = H ijkl ( e i e l ) ( e j e k ) = H ikki
152
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
X 12 ( T ) = X 12 ( T ij e i e j ) = T ij e j e i = T T
2
2
X 1 = X 1 ( ijk e i e j e k ) = ijk e j e i e k
X 13 = X 13 ( ijk e i e j e k ) = ijk e k e j e i
(19.20)
(19.21)
(19.22)
The contraction operator and the swap operator are almost never mentioned explicitly in publications but they are nonetheless very useful in generating theorems because they are both linear operations. Namely,
C IJ ( A + B ) = C IJ ( A ) + C IJ ( B ) for all scalars
and for all tensors A and B of arbitrary (but equal) order (>2).
(19.23)
(19.24)
(19.25)
45 ( u v )
u v = C 23
(19.26)
and
153
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
on
Brann
fifth order tensor having ijklm components given by ijk u l v m . The contrac45 reduces the order from five down to one by contracting the
tion operator C 23
2nd index with the 4th index and the 3rd index with the 5th index to obtain
ijk u j v k , which is indeed the indicial expression for the i th component of
u v . It may appear that we have taken a fairly simple operation (the cross
product)
and written it in an insanely complicated manner.
Keep in mind: we are not advocating the use of the contraction and swap
operations in published analyses. We merely want to drive home the point
that virtually all tensor operations can be expressed using contraction and
swap operators. Contractions and swaps are particularly useful when applied
to tensor-integral calculus. Specifically, suppose that f ( x ) denotes some n th
order tensor valued function of a vector x . If C and X denote
any contraction
C f ( x ) dV = C [ f ( x ) ] dV
(19.27)
(19.28)
X f ( x ) dV = X [ f ( x ) ] dV
n (
)d A =
) dV
(19.29)
Here, the notation n ( ) denotes the outward unit normal n to the integra some quantity ( ) of arbitrary tensororder, and ( )
tion surface multiplied
denotes the gradient of that tensor.
Because the contraction operation commutes with integration, one can
immediately write down the classic divergence theorem by taking the C 12
contraction of both sides of Eq. (19.29) to obtain
n (
)d A =
) dV
(19.30)
Another corollary is obtained by inserting the cross product into Eq. (19.29),
which we know we can do because we know that a cross product is expressible
in terms of contractions (which in turn commute with integration). Thus,
n (
)d A =
) dV
(19.31)
154
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
Incidentally, despite the presence of the cross product, this theorem is not
equivalent to Stokes theorem.
Consider a particular indicial form of Eq. (19.29) in which the operand is a
second-order tensor:
ni T jk d A
xi T jk dV
(19.32)
2
C 1 ( T ) = C :T = C ij T ij
(19.33)
We know from Eq. (19.17) that
C 12 ( G ) = trT = T kk = ij T ij
(19.34)
Comparing the above two equations, we note that C is simply the identity
as the identity tensor
tensor. Hence, the contraction operation can be viewed
double dotted into two specified base vectors. This point of view is useful when
considering differentiation. The identity is a constant tensor and its derivative
is therefore zero. It is often useful conceptually utilize the contraction in gradient identities. For example,
(19.35)
w = I : w
w dV
I : ( w ) dV
= I : w d V
(19.36)
These formulas show that knowing the complete gradient tensor w is really
the most useful. You can just take the trace of this total gradient tensor to
immediately obtain the divergence. Likewise, you can take the (negative)
axial part of the total gradient tensor to obtain the curl:
axial w = : w = w
cca B
ranno
(19.37)
The swap tensor. Recall from Eq. (19.20) that the effect of the swap operation
on a second-order tensor T is to simply produce the transpose of that tensor:
155
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
X 12 ( T ) = T T
(19.38)
(19.39)
T T = S :T
Writing this out in indicial notation reveals that
S = in jm e i e j e m e n
(19.40)
156
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
we are implicitly telling the reader that the vector y can be expressed as a
(20.2)
cca B
ranno
(20.3)
d y is linear
with respect to d x . Thus,
exists
a second order tensor, which
we will denote d y d x such that
d
y
d y = ------- d x
(20.4)
d x
* In Eq. (20.2) we have written y i x j on the reasonable assumption that our readers will understand this to mean the partial derivative with respect to x j , holding all other components of x con
stant. Except in this very special case, we strongly recommend explicitly showing what variables
are held constant in partial derivatives whenever there is even the slightest chance of misinterpretation or confusion by your readers. This issue pops up continually in thermodynamics. The partial
derivative of pressure P with respect to volume V takes on different values depending on the conditions at which the measurement is made. If, for example, the variation of pressure with volume is
measured at constant temperature T , then the derivative should be written ( P V ) T . However, if
the measurement is made under reversible insulated (no heat flow) conditions, then the derivative
should be written ( P V ) s , where s is entropy. The two derivatives arent equal!
157
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
(20.5)
ASIDE #1: total and partial derivative notation. Note that we have
quietly introduced our notational preference to use a d when writing d y d x
the
even though we must, by necessity, use a in the partial derivative on
right-hand-side. This choice has been made quite deliberately in order to maximize the analogies that tensor calculus has with ordinary scalar calculus. To
get a better appreciation of why we have made this choice, consider the scalar
calculus situation in which some variable y is expressible solely as a function
of an independent variable x . Then, of course, you would naturally write
dy dx , not y x . The very act of writing dy dx tells the reader that y
really can be written solely in terms of x . If it turns out that y can be alternatively written as a function of two other variables, u and v, then we could use
the chain rule to write
y du
y dv
dy
------- = ------ ------- + ------ ------ u v dx v u dx
dx
(20.6)
Note that we use d on the left-hand-side, and on the right. Our situation
with vector derivatives is very similar. We have a vector y that can be computed if we know the single vector x . We say this object is a single vector
Note the distinctions in the use of d and in this equation. Its not hard to
show that
dx j
--------- = e j
(20.8)
dx
Likewise, since the base vectors themselves are fixed when using rectangular
Cartesian coordinates,
( yi ei )
yi
y
- = --------e
------- = ----------------x j i
x j
x j
(20.9)
Putting Eqs. (20.8) and (20.9) into (20.7), taking care to preserve the order of
dyadic multiplication of the base vectors, gives us back our result of Eq. (20.5).
158
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
There are other analogs with conventional scalar calculus that motivate
our use of d on the left hand side of Eq. (20.5). In scalar calculus, we know
that
dx
1
------- = -----------------dy
dy dx
(20.10)
(20.12)
Equality holds only if both partial derivatives hold the same variable constant
in the derivative (in which case we are implicitly restricting our attention to a
world in which that variable is always constant, and in such a world, y
would be expressible solely as a function of x ). To convince yourself that
Eq. (20.12) is correct, consider polar coordinates in which x and y are expressible as functions of r and . Namely
x = r cos
y = r sin
and
(20.13)
x
---- r
y
------
r
y
---- r
cos r sin
sin r cos
(20.14)
r
---- y x
------
x y
------
y x
x
------
r
x
---- r
y
------
r
y
---- r
cos r sin
sin r cos
cos sin
= sin cos
-------------- -----------r
r
(20.15)
cca B
ranno
(20.16)
159
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
r
----- = cos
x y
(20.17)
Therefore,
r
1
----- ---------------------- x y ( x r )
(20.18)
If (for some strange reason) we wished to know the variation in r with respect
to x holding constant, then we would be able to write
r
1
----- = ---------------------- x
( x r )
(20.19)
The only distinction between the above two equations is what variable is
being held constant. In one case equality holds; in the other it doesnt thats
why it is so essential to explicitly show whats being held constant unless you
are 100% sure there is no chance of confusion. Dont be so sure. In classic multivariable, scalar, calculus, the very act of writing y x tells your reader that
y can not be written solely in terms of x it must be expressible in terms of
x and whatever variable is being held constant in the partial derivative.
Thats why we do not use . When we write d y d x , we want our readers to
things, and we would be diligent to show what variable is held constant in the
derivative if there is any chance for misinterpretation of what this second
variable is.
ASIDE #2: Right and left gradient operations. Note that the ij component
in Eq. (20.5) is y i x j . Some books define the gradient such that the ij component is y j x i . When they do this, they are required to redefine their definition of the vector-vector gradient to be
dy
d y = d x -------
(20.20)
d x
This is fine, although we frown at instances where the gradient conventions are mixed inconsistently.
The nabla or del gradient operator. Researchers who use y j x i
typically do so because this convention permits them to use the alternative
notation y , where the nabla or del operator is heuristically defined such
that
( )
( ) = ----------- e i ,
x i
(20.21)
160
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
y
d y = d x ( y ) , which implies that ( y ) ij = ---------j
(20.22)
x i
(20.23)
d y d x = y
(20.24)
Then the ij component of y would be heuristically similar to the ij compo be defined such that
nents of y b . Thus, y would
y
d y = ( y ) d x , which implies that ( y ) ij = --------i
(20.25)
x j
Comparing this with Eq. (20.22) shows that the two types of del operators
are related by a simple transpose
( y ) = ( y )T
(20.26)
Both the backward and forward operating del are used by Malvern and we
also find that they are both useful, depending on the application at hand.
However, for the applications in this report, a del operator is never needed.
We will always use the fraction-like notation of Eq. (20.5), which is technically a left-operating derivative. In general, the indices are placed in order
first on the dependent variable y i and then on the independent variable x j .
Since we will always follow this convention, we may write d y d x as
dy
------(20.27)
dx
without ambiguity.
Our convention of placing indices first on the dependent variable and then
on the independent variable extends to define higher order derivatives as well.
For example, the notation
dU
-------dY
cca B
ranno
(20.28)
161
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
U ij
------------------e e e e e e
Y klmn i j k l m n
(20.29)
(20.30)
where the superscript 1 is the tensor inverse. Of course, this property holds
only if the inverse exists (i.e., only if det ( d y d x ) 0 ), just as Eq. (20.30) holds
only if dy dx 0 .
In contrast to Eq. (20.27), when we write a derivative with the symbol
instead of the d symbol, we are telling the reader that the dependent variable depends on more than one tensor or vector quantity. For example, by
writing the notation
y
-----(20.32)
x z
we are implicitly telling the reader that y can be written as a proper function
of x and z . The subscript also indicates that the vector z is being held con in the
derivative. Thus, the above derivative quantifies
how the vector y
stant
As before, if y is a function of x and z , then y i depends on the x j and z k
components. Hence,
* The situation is like having y= f ( x ) and x = g ( u, v ) . By the chain rule,
y
y
dy = ------ du + ------ dv . On the left side, we use d, while on the right side we use .
v u
u v
162
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
y
y
------ = --------i e i e j .
x z
x j z
(20.34)
y 1
x
------
= -------
(20.35)
y z
x z
This is true because the same quantity namely, z is being held constant
x w
y z
in general
(20.36)
Okay, if the above relation does not hold, does anything LIKE IT hold?
If you really wish to relate ( x y ) z to ( y x ) w , then you must be extra
seeking
careful with your logic. The basic motivation
for
such a relationship in
the first place is that you must not have the actual function for x as a function
some other
of y and z available. Instead, you have y as a function of x and
variable
w . Recall that the very act of writing
( x y ) z tells your reader that
function
of x and z . Now, the very act of writing the expression ( y x ) w ,
implies that y must be expressible as a function of x and some other variable
w . Thus, we have simultaneous implied functions
y = y ( x, z ) and y = y ( x, w )
(20.37)
The only way these can both be true is if
w = w ( x, z )
Now we can apply the chain rule to write
y
y
y
x
w
------ = ------- ------- + ------- ------
x z
x w x z w x x z
or, noting that ( x x ) z is just the identity tensor,
y
y
w
------- = ------- + ------- ------ x z
x w w
x z
x
Putting this into Eq. (20.35) gives
y
y
x
w
------ = ------- + ------
- ------ y z
x w w
x z
x
cca B
ranno
(20.38)
(20.39)
(20.40)
(20.41)
163
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
The parameter s is here assumed to vary freely from to . The non-parametric equation for this space curve is simply x 2 = e x1 and x 3 = 0 .
The derivative of x with respect to s is
dx
------- = E 1 + e s E 2
ds
(20.43)
Because the curve is one-dimensional, we know that each position x corresponds to a unique value of the parameter s . Conversely, each value of s is
associated with exactly one position vector x . Hence, s may be regarded as a
(20.44)
(20.45)
Both of these expressions will give the correct value for s as long as x lies on
the space curve. However, directly differentiating s with respect to x gives
E2
ds
Eq. (20.45) gives ------- = -----(20.47)
d x
x2
These two expressions are perpendicular to each other, so they cannot possibly
be equal. The discrepancy must be resolved by using a projection of the increment d x onto the space of allowable increments. From Eq. (20.42), note that
d x = ( E 1 + e s E 2 ) ds
(20.48)
We can define the projector onto this vector as
(20.49)
164
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
E1 E1 +
E +
E +
E
21
12
2 2P = ------------(20.50)
- = ----------------------------------------------------------------------------------------------
1 + e 2s
If d x is an arbitrary increment in x that may or may not lie on the space
we can write the set of all allowable
curve,
increments in the form
es E
es E
cca B
ranno
e 2s E
( d x ) = P d x
(20.51)
ds
ds
ds = ------- ( d x ) = ------- P d x
(20.52)
d x
d x
from which it follows that the proper expression for the derivative of s with
respect to allowable increments in position must be given by
ds
ds
------- = ------- P
d x
dx
------- = ------------------------dx
1 + e 2s
(20.53)
(20.54)
E1 + e s E2
ds
------- = ------------------------(20.55)
dx
1 + e 2s
The above two results are identical and should be used whenever the derivative of s with respect to x is needed. Using this expression for the derivative
Substituting Eqs. (20.53) and (20.54) into the right hand side leads to
(20.56)
dx
------- = P
(20.57)
dx
P ( d x ) = ( d x )
(20.58)
Thus, the projection operator is the identity tensor along the space curve!
165
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
pq = E pqmn mn
(20.60)
(20.61)
Because the strain tensor is symmetric, the tangent stiffness tensor can (and
should) be replaced by
pq
sym
E pqmn = ------------ P ijmn
ij
(20.62)
166
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
(20.63)
where
f o f ( x o, y o )
f ( x, y )
f xo = --------------------- evaluated at ( x, y ) =
x
f ( x, y )
f oy = --------------------- evaluated at ( x, y ) =
y
2 f ( x, y )
o =
f xx
----------------------- evaluated at ( x, y )
x 2
2 f ( x, y )
f oyy = ----------------------- evaluated at ( x, y )
y2
2 f ( x, y )
o = ----------------------f xy
- evaluated at ( x, y )
xy
2 f ( x, y )
o = ----------------------- evaluated at ( x, y )
f xy
xy
( x o, y o )
( x o, y o )
= ( x o, y o )
= ( x o, y o )
= ( x o, y o )
= ( x o, y o )
f oyx
cca B
ranno
(20.64)
1
f ( x ) = f o + f o ( x x o ) + ----- ( x x o ) f ( x x o ) +
(20.65)
2!
where
167
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
f o f ( xo )
df ( x )
f
f o = -------------- - = -------- e i evaluated at x = x o
x
2
2
d f ( x)
f
f = ----------------(20.66)
- = ----------------- e i e j evaluated at x = x o
x i x j
d xd x
o
In Eq. (20.65) the first order derivative f o is a vector and it appears in a vec
tor inner product with ( x x o ) . The second-order
derivative f is a second
o is the same
order tensor, and it is dotted from both sides by ( x x o ) , which
thing as taking the second-order tensor inner product
f : [ ( x x o ) ( x x o ) ]
(20.67)
Following along with this pattern, the next term in Eq. (20.65) would be
1
----- f [ ( x x o ) ( x x o ) ( x x o ) ] ,
(20.68)
3! o
where would denote the third-order tensor inner product* so that the component form of the above expression would be
1 3 f
----- --------------------------3! x i x j x k
[ ( x i x io ) ( x j x oj ) ( x k x ko ) ]
(20.69)
xo
In this case, the component form is more clear than the direct notation form,
so it might be a better choice in publications, especially given that the triple
dot notation is highly nonstandard and might only confuse your readers.
* Clearly, indicating the order of the inner product by the number of stacked dots could lead to
some very weird notation for higher order tensors. We have seen the notation :: used for the fourthorder inner product. A circled number can always be used for higher order inner products.
168
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
cca B
ranno
169
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
170
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
Rebe
REFERENCES
1Aris,
R., Vectors, Tensors, and the Basic Equations of Fluid Machanics. Englewood Cliffs, N.J.: Prentice-Hall, 1962.
2
this.
3Budiansky,
Bernard, (1990) Tensors, in: Handbook of Applied Mathematics 2nd Ed, Edited by Carl E. Pearson. Van Nostrand Reinhold.
4Brannon,
R. M., http://me.unm.edu/~rmbrann/curvilinear.pdf
5Brannon,
R. M., http://me.unm.edu/~rmbrann/Rotation.pdf
6Brannon,
R. M., http://me.unm.edu/~rmbrann/RadialReturn.pdf
7Brannon,
R. M., http://me.unm.edu/~rmbrann/MohrsCircle.pdf
8Brannon,
C. Ray, and Barrett, Louis C., Advanced Engineering Mathematics, McGraw-Hill (1982).
10Brannon,
cca B
ranno
Wiley (1979).
12
Malvern, Lawrence E., Introduction to the Mechanics of a Continuous Medium, Prentice-Hall (1969).
This is one of the best known references on elementary continuum mechanics. HomeLibrary
13Smart,
G.F., Smith, M.M., and Rivlin, R.S., Integrity bases for a symmetric tensor and a vector The crystal classes. Arch. Rational Mech. Anal.,
Vol 12 (1963).
The title says it all. This paper shows the most general quadratic form that can be constructed
from the isotropic invariants for a crystal class. 970806
171
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
15
Spencer, A.J.M., Theory of Invariants, in Continuum Physics I, edited by A.C. Eringen Academic, New York (1971).
Derives the integrity bases for arbitrary numbers of vectors and tensors for (i) full orthogonal
group (ii) proper orthogonal (rotation) group, (iii) transverse isotropy, and (iv) crystal classes.
16
Varadan, Vasundara V.; Jeng, Jiann-Hwa; and Varadan, Vijay K., Form
invariant constitutive relations for transversely isotropic piezoelectric materials. J Acoust. Soc. Am. 82 (1) pp. 337-342 (1987).
This paper shows how to derive transversely isotropic constitutive relations using integrity
basis for the transverse orthogonal group of transformations. 970722A
17Continuum
mechanics textbook
from my class.
18Tai,
NY (1994).
20Chadwick
21Stewart,
24
Horn, Roger A., and Johnson, Charles R., Matrix Analysis, Cambridge University Press, Melbourne Australia (1985).
25Mathematica
reference
26
27
28Ramkrishna,
172
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
DRAF
30
Rebe
cca B
ranno
31Boccara,
Rebeccas math handbook that contains the discussion of tensors in terms of dyads.
173
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
T
F
A
R
D
cca
Rebe
Brann
on
174
Copyright is reserved. Individual copies may be made for personal use. No part of this document may be reproduced for profit.
INDEX
A
adjoint 19
adjoint, not the same thing as adjugate 81
adjugate 19
affine 6
alternating (permutation) symbol vs. alternating TENSOR 40
alternating symbol (also called the permutation symbol or
Levi-Civita density) 11
alternating symbol, cyclic property 11
alternating TENSOR 40
alternating tensor 42
angle between two vectors 35
area, computed using cross products 39
axial tensor 42
axial vectors 42
axial vectors, applied to rotation 93
B
basis 6
basis transformations 104
basis, effect of changing 8
basis, for a tensor 60
basis, style of notation 73
binary operation 6
C
calculus 157
Cayley-Hamilton theorem 122
characteristic invariant 17
closed 4
codomain 4
cofactor 15
COFACTOR tensor 80
cofactor-determinant connection 19
complementary projectors 85
completing the square, generalized to vectors/tensors 133
component, of a tensor 71
component, of a vector in direction of another vector 38
components of a vector 25
components, of a vector 36
components, of a vector with respect to NON-orthonormal
basis 37
composition (dot product) between two tensors 78
composition of two tensors 71
constraints 5
Continuous functions are vectors 147
contraction operator 151
contraction tensor 155
coordinate transformations 104
Cramers rule for finding the inverse 82
cross product 39
cross product, associated tensor 91
cross product, between orthonormal base vectors 39
cross product, between two tensors 100
cross product, handedness issues 40
curvilinear 6
cyclic property of the alternating symbol 11
cylinder, equation for 132
D
definition of a matrix 7
definition of a vector 24
definition of a vector, generalized 146
definition of the 3x3 determinant 16
definition of the cofactor 20
definition of the inverse 20
definition of the laboratory basis 25
definition of the permutation symbol 40
derivative, of a vector with respect to a vector 157
Determinant 14
determinant, definition 16
determinant, derivative of 83
determinate, using cofactors 19
Deviatoric tensors 89
deviatoric tensors 97
diagonalizable tensors 125
difference between a matrix and a tensor 112
difference between a tensor and a matrix 108
dimension 5
dimension, of a matrix 7
direct notation 110
direct, style of notation 73
direction cosines 105
domain 4
dot product, between a tensor and a vector 69
dot product, between a vector and a tensor 70
dot product, between orthonormal base vectors 36
dot product, between two tensors 71
dot product, between vectors 35
dot product, dotting a tensor from both sides 71
double dot product 98
double-bar (vector/tensor magnitude) 101
double-dot product 98
dummy index 26
dyad "no-symbol" notation 62
dyad, component matrix 63
dyad, scalar multiplication rule 64
dyadic 66
dyadic multiplication 61
dyadic sums, reducibility 64
DYADS 61
dyads, REDUCIBILITY 65
dyads, sum of two 64
E
e-delta identity 11
e-delta identity, easier mnemonic 12
e-delta identity, fast way 12
e-delta identity, multiple summed indices 13
e-delta identity, three summed indices 14
e-delta identity, two summed indices 14
Eigenprojectors 126
Eigenvalue problems 122
eigenvalues, algebraic vs. geometric multiplicity 123
eigenvector, right and left 122
Einstein, quote 1
ellipsoid, equation of 130
Equation for a cylinder 132
Equation for a sphere 130
Equation for an ellipsoid 130
Equation of a line 129
Equation of a plane 128
Euler angles 94
even functions 37
F
free index 11, 26
functional analysis 3
G
Glide plane 43
gradient 157
gradients, direct notation 75
H
handedness 41
homogeneous functions 37
hyperboloid, equation of 132
hyperplanes 6
I
idempotent 53
identity matrix 10
identity tensor 77
implicit sums as independent variables 29
improper orthogonal tensors 90
indexing, aesthetic/courteous 31
Indicial notation in derivatives 29
indicial, style of notation 73
inertia tensor 114
Inner product spaces 147
inner product, between tensors, identities 101
inner product, of two column matrices 9
invariance 108
invariants, characteristic 118
invariants, more than three 118
Invariants, of a sum of two tensors 121
invariants, primitive 117
invariants, scalar 116
invariants, traces of powers 118
Inverse 19
inverse, Cramers rule 82
inverse, of a function/transformation 5
Inverse, of a tensor plus a dyad 122
inverse, of a tensor plus a dyad 83
irregular 37
Isotropic fourth-order tensors 143
isotropic materials 140
Isotropic second-order tensors, in 2D 142
Isotropic second-order tensors, in 3D 140
isotropic tensors 97
K
Kronecker delta 10
Kronecker delta, index-changing property 33
Kronecker delta, summing 34
L
laboratory base vectors 24
laboratory basis 24, 25
laboratory triad 24
left-handed 41
Levi-Civita density (also called the permutation symbol or
alternating symbol) 11
line, equation of 129
LINEAR 58
linear 4
linear combination 5
linear manifold 6
linear operators/transformations 58
linear space 6
linear transformations, existence of an associated tensor 76
linear transformations, finding the associated tensor 76
linearly independent 5
M
magnitude of a tensor 101
magnitude of a vector 101
Material symmetry 140
matrices, not the same as tensors 112
matrix vs. tensor 108
matrix, definition 7
matrix, dimensions 7
matrix, product 8
minor 15
minor, signed 15
Mohrs circle (footnote) 123
N
nested submatrices 17
notation
three styles (direct, basis, indicial) 73
notation, under-tilde 34
notation, WHY FRET OVER IT? 79
null space 4
O
odd functions 37
one-to-one 5
operators, linear 58
order of a tensor 102
order vs. rank 102
orientation of a vector 38
orthogonal tensors 89
orthogonal tensors, improper 90
orthogonal tensors, proper 90
orthogonal, improper 90
orthogonal, projections 46
outer product, of two column matrices 10
P
permutation (alternating) symbol, handedness issues 40
permutation symbol (also called the alternating symbol or
Levi-Civita density) 11
permutation symbol, cyclic property 11
plane, equation of 128
polar decomposition 135
polar decomposition, as a nonlinear projection 138
polar decomposition, fast method in 2D 139
principal directions 122
principal minors 17
principal submatrix 17
projection decomposition, finding magnitudes 101
projection of a vector onto another vector 38
projection theorem 55
projection, and subspaces 151
projection, of a vector onto another vector 44
projection, to a desired target space 85
PROJECTIONS 44
projections, fourth-order oblique 102
projections, generalized 53
projections, linear 53
projections, nonlinear 53
projections, orthogonal linear 44
projections, orthogonal, basis interpretation 47