You are on page 1of 15

Contents

1 Some Actors and Acts 3


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Some Actors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Some Acts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 The Integral Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Divergence of Vector (and Tensor) Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Rotation or Curl of Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 The Dirac Delta Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

1
2 CONTENTS
Chapter 1

Actors and Acts Used in Continuum


Physics

1.1 Introduction
We will be developing the laws of continuum physics throughout the first portion of this class. We will do so
in an uncommon but highly pedagogic way by starting with the laws that describe the discrete movement
of individual atoms and then averaging over the molecular dynamics. The emergent continuum laws so
obtained come in the form of partial differential equations (PDEs) that determine how fields are changing
in time at each point in space based on how the fields are varying in space in the immediate neighborhood
of that point. The second portion of the class will be devoted, for the most part, to math techniques for
solving the PDEs (with a heavy dose of Fourier analysis). You can think of these two portions of the class as
the equivalent of climbing a mountain. The first part of the class, in which we develop the continuum laws,
is akin to the hard slog as you work your way up the mountain until you are at the peak and all the laws
needed to describe any Earth response are available and understood to you. The second part of the class in
which we solve the PDEs is akin to coming down the mountain as we choose from many different trails (the
various math techniques) that get us to where we want to go. Like climbing a mountain, you will find the
first part of the class to be the most difficult and the descent to be relatively easier.
The various types of spatial derivatives (gradient, divergence, curl) that characterize the ways the fields
are varying in space are reviewed in this first chapter that is largely devoted to ideas from vector calculus
(sometimes called multivariable calculus). However, this first chapter is not purely review. From the very
beginning we will be introducing the idea of tensors which are key players (or actors) in continuum physics.
For some reason, tensors are not usually introduced to undergraduates in the United States even if, for
example, you cannot properly describe the way a seismic wave propagates without them. Because tensors
are not usually introduced to US undergraduates, the laws of continuum physics that we use to quantify
nearly all Earth processes are also not commonly taught to undergraduates. However, there is no reason
to hide you from tensors or the laws of continuum physics. This class will bring together all the math and
physics you have learned up to now and package it in the form that Earth scientists (e.g., geophysicists,
hydrogeologists and climate scientists) actually use in practice. You will have more confidence going to
graduate school and looking at journal articles if you are comfortable with the material in this class.
This class would ideally be extended into two classes EPS 104A and EPS 104B. In EPS 104A, which
corresponds to the present class, the focus is on analytical understanding of the physics and math and involves
pencil and paper work. The goal is to build intuition and hands on familiarity with the ideas. Simulating
actual Earth response is called the forward problem and is usually performed numerically due to the
complicated nature of the Earth. Recording data of various types and minimizing the difference between
the recorded data and simulations of the data with the goal of determining the properties throughout the
Earth is called the inverse problem and is also a numerical exercise in nearly all cases. EPS 104B, which
does not exist, would therefore be about how to perform numerically various types of forward and inverse
problems. You will have to gain programming and numerical-algorithm experience elsewhere.

3
4 CHAPTER 1. SOME ACTORS AND ACTS

1.2 Some Actors


Fields: Any physical quantity continously distributed over the space of some region is called a field.
Fields can be scalars, vectors, or tensors.

Scalars: A field quantity that has no intrinsic direction is called a scalar field. Examples: temperature,
pressure, etc. In the nomenclature of tensors, a scalar can be called a zeroth-order tensor.

Vectors A field quantity that has a direction associated with it is called a vector field. Examples:
electric fields, fluid velocity, etc. Vector fields are represented at each point by an arrow whose length
denotes the amplitude of the field at that point. In the nomenclature of tensors, a vector can be called
a first-order tensor. Vectors are written analytically in different ways:

r=
b vector used to define points in space
= x1 x1 + x2 x2 + x3 x3 = x x
+ y y + z z (1.1)
i
= (x1 , x2 , x3 ) = xi x
(summation over repeated indices always assumed)

a(r) = a(x1 , x2 , x3 ) =
b vector field defined at each point r
= a1 x1 + a2 x2 + a3 x3 = ax x
+ ay y + az z (1.2)
i
= (a1 , a2 , a3 ) = ai x

IMPORTANT: Whenever an index appears twice in an expression, you always sum over that
index. The index that is summed over is sometimes called a dummy index because the index
does not survive
P3the summation and could be given any name. For example, we have ai bi =
aj bj = an bn = n=1 an bn = a1 b1 + a2 b2 + a3 b3 where the i, j and n are the dummy indices that
we sum over. The summation over repeated indices in vectorial and tensorial expressions is called
the Einstein summation convention and simply saves us from having to write the summation sign
over and over.

Another type of vector is the vector operator that we call the gradient operator that is defined

=x
+ y + z =
b gradient operator (a vector operator) (1.3)
x y z

For example, if (r) = (x, y, z) is some scalar field then the gradient of is = x
x + y y + z z
and is a vector that we can also write = x i /xi using the summation convention. The gradient
vector is oriented in the direction that the field is increasing the most rapdily and the amplitude
|| gives the rate of that maximum increase.
For a vector a = ai x i we call the ai the components of the vector and call the unit vectors x
i in each
direction i the base vectors. Note that a vector at some point in space is an arrow with a length and is
completely independent of the coordinate system we use to describe it. So a happily exists as the same
arrow and does not change if we rotate the coordinate system. Note, however, that the components of
the vector ai will change as we rotate our coordinate system (alter the orientations of the base vectors)
or switch to another coordinate system such as cylindrical coordinates (redefine the base vectors).
Finally, note that some authors put an arrow above a symbol to denote that it is a vector field; i.e., ~a.
When working in typed text, we always use a bold-face symbol to denote a vector. When writing by
hand, we have elected not to use an arrow over a symbol but instead use a squiggly underscore; i.e., a.
e

2nd-Order Tensors: A field quantity that acts as the proportionality between two vector fields at
a point in space is called a second-order tensor. Sometimes people just call these quantities tensors.
Another word that is synonymous to tensor is dyad. We write a second-order tensor as
1.2. SOME ACTORS 5

T (r) = a tensor field defined at each point r


x
= Txx x + Txy x
y + Txz x
z
+ Tyx yx
+ Tyy yy + Tyz yz (1.4)
+ Tzx zx
+ Tzy zy + Tzz zz
ix
= Tij x j (summation over repeated indices assumed) (1.5)

Txx Txy Txz
= Tyx Tyy Tyz (1.6)
Tzx Tzy Tzz
+ ay y + az z is the sum of three vectors in the three coordinate directions,
Just like the vector a = ax x
so the second-order tensor T is the sum of 9 second-order tensors as made explicit in Eq. (1.4).
We see that a second-order tensor has much in common with a matrix. Much of what you learned about
matrices in linear algebra applies to how we use second-order tensors. The main difference between
a matrix and a second-order tensor is that although a matrix may have any dimension (N M )
and correspond to any proportionality between an M and N dimensioned vector array, a second-order
tensor is a field quantity distributed through three-dimensional space and is always a (33) matrix and
is a physical field that is the proportionality between two vector fields that have a physical meaning.
Other examples of tensors include two vector fields that sit side by side to each other in an expression
without a scalar or vector product (that are defined in a later section) between them:
+ ay y + az z)
ab = (ax x (bx x
+ by y + bz z)


= ax bx xx + ax by xy + ax bz xz (1.7)
+ ay bx yx
+ ay by yy + ay bz yz
+ az bx zx
+ az by zy + az bz zz.

Another example of a second-order tensor is the gradient of a vector field written as
 

a = x + y + z + ay y + az z)
(ax x (1.8)
x y z
ax ay aj
= x
x + y + . . . =
x xi xj
x x xi
again with summation over the repeated indices as always.
ix
Just like with a matrix, we can talk about the transpose of a 2nd-order tensor T = Tij x j and write
TT = j x
b the transpose of T = Tij x i = Tji x
ix
j . (1.9)
Thus to perform the transpose we can either flip the indices on the components Tij Tji of the tensor
or flip the position of the base vectors as they sit side by side.
Note that like with a vector, a tensor T exists at a point and is independent of the coordinate system.
If we rotate or change coordinate systems, T does not change. However, the components of the tensor
Tij will change as we change the coordinates because the base vectors x i are changing. When working
in typed text, we always denote a tensor with bold type. When we write a second-order tensor by
hand, we use two squiggly underscores T = T .
e
e

Higher-order Tensors: The generalization to higher-order tensors is straightforward. A third-order


tensor is written
ix
3 P = Pijk x j x
k (1.10)
a four-order tensor as
4Q ix
= Qijkl x j x
k x
l (1.11)
and so on for still higher-order tensors. Summation over each index that appears twice is again assumed.
We will not have occasion to consider tensors higher than the 4th order. We write an nth-order tensor
by hand as n Q for n > 2.
e
6 CHAPTER 1. SOME ACTORS AND ACTS

1.3 Some Acts


Scalar Products: A scalar product between two vectors is the product of the amplitude of the two
vectors after one of the two vectors is projected into the direction of the other vector. We have:
q
a b = |a||b| cos where |a| = a2x + a2y + a2z . (1.12)

So a b = 0 if a b and x
y = 0 and x
z = 0, but x
x
= 1, etc. Using
these rules, we thus have

a b = [ax x [bx x
+ ay y + az z] + by y + bz z]

= ax bx + ay by + az bz (1.13)
= ai bi .

The scalar product is also called the dot product or the inner product.
What if vector a is related to vector b at some point in space? How do you obtain a given b? That
is what tensors do. Working in two spatial dimensions so that we have less components to write out
(with implied summation over repeated indices as always)

a=T b (1.14)
 
Txx xx
+ Txy x y
= (bx x
+ by y)

+ Tyx yx
+ Tyy yy
+ (Tyx bx + Tyy by ) y
= (Txx bx + Txy by ) x

or using the familiar matrix multiplication for the scalar product


    
ax Txx Txy bx
= . (1.15)
ay Tyx Tyy by

IMPORTANT: Second-order tensors are always maps between two vectors that are physically
related to each other at point r. You cannot visualize a second-order tensor (or higher-order
tensors). But you can picture in your minds eye the two vectors (arrows) that are related to
each other and thus imagine the mapping (tensor) that takes the one vector to the other.

We can also speak of the double dot product : between tensors. So for example

ab : cd = (a d)(b c) (1.16)
i dj x
= (ai x k cl x
j ) (bk x l)
i x
= ai dj bk cl (x k x
j ) (x l) . (1.17)

Because of the nature of the scalar product, we thus have l = k and j = i. As such we end up with

ab : cd = ai bk ck di with summation over repeated indices (1.18)


= a1 b1 c1 d1 + a2 b1 c1 d2 + a1 b2 c2 d1 + a2 b2 c2 d2 in 2D. (1.19)

Further,

ix
S : T = Sij x j : Tkl xk xl (1.20)
= Sij Tkl (xi x k x
j ) (x l) (1.21)
= Sik Tki with summation over repeated indices.

Note that for two 2nd-order tensors S and T , we have that S T 6= T S. For the double dot product
however, we do have S : T = T : S.
1.3. SOME ACTS 7

IMPORTANT: In tensorial expressions that involve multiple tensors or vectors, use different
indices to represent each tensor and/or vector. You can then pull out all the tensor or vector
components to the left of the expression and leave all the base vectors behind. However, you
always need to leave the base vectors in the order that they appear and perfom the scalar products
while respecting this order.

ix
The identity tensor I is defined I = ij x j where ij is called the Kroneker delta function and is
defined
(
0 if i 6= j
ij = (1.22)
1 if i = j.

So upon summing over the indices we have I = x 1x


1 + x2x
2 + x
3x
3 = x
x + yy + zz.
The identity
tensor works as follows: A I = I A = A for any tensor A.

Finally, a double-dot product with the identity tensor results in A : I = Aij ji = Aii = tr(A) which
is called the trace of tensor A. The trace is the sum of the tensor along the diagonal.

Vector Products: A vector product between two vectors is a vector that is perpendicular to the two
vectors and that has an amplitude equal to the area of the parallelogram formed with the two vectors
as sides. We have:

c = a b; c to both a and b (1.23)


|c| = |a||b| sin (1.24)
ab=0 if a||b. (1.25)

Use the right-hand rule to determine the sense of a b. Note that a b = b a.


We can thus obtain

y = z;
x z = y;
x x
x =0 (1.26)

and so forth for all the vector-products between all base vectors. Using the above,
we can write

a b = (ax x
+ ay y + az z) (bx x + by y + bz z)

= ax by z ax bz y ay bx z + ay bz x + az bx y az by x
(1.27)
= (ay bz az by ) x
(ax bz az bx ) y + (ax by ay bx ) z.

We can thus write the vector product using the matrix determinant in the following way

y z
x
a b = ax ay az where | | denotes taking the determinant. (1.28)
bx by bz

Note that the vector product is also called the cross product.

Doing Cross (Vector) Products with Dot (Scalar) Products: It is convenient for proving
identities involving the cross product to write a cross product in a way that only involves dot products.
To do so, we introduce the alternating or Levi-Civita third-order tensor 3  defined as

3 ix
= ijk x j x
k (1.29)
8 CHAPTER 1. SOME ACTORS AND ACTS

where


+1 for clockwise indices: 132, 321, 213

ijk = 1 for counterclockwise: 123, 231, 312

0 for every other combination.

To perform the cross product with the Levi-Civita tensor we do the following
a b = 3  : ab = ijk xi x
j x k : al x
l bm x
m
= ijk al bm x k x
i (x j x
l ) (x m)

= i
ijk ak bj x (1.30)

1 [a2 b3 a3 b2 ] + x
=x 2 [a3 b1 a1 b3 ] + x
3 [a1 b2 a2 b1 ]
which is identical to what we obtained earlier.

1.4 The Integral Theorems


Fundamental Integral Theorem of 3D Calculus : For some finite volumetric region bounded
by the closed surface that has an outward normal n at each point of , we have

Z Z
(r) d3 r = n(r) d2 r Fundamental Integral Theorem

(1.31)

where is a scalar, vector, or tensor field of any order. We will not pause to prove this most important
integral theorem that is used repeatedly in physics and that can be considered a 3D generalization of
the fundamental theorem of calculus
Z b
d(x)
dx = (b) (a). (1.32)
a dx
Note that we are using the following notation for the infinitesmal volume and surface elements in the
integrals of Eq. (1.31)
d3 r = dx dy dz = dV = volume element
d2 r = dS = surface element.
Using the fundamental integral theorem, it is straightforward to obtain the divergence or Gausss
theorem Z Z
3
ad r = n a d2 r. Divergence Theorem (1.33)

In the next section, we will use the divergence theorem to interpret the physical meaning of the
divergence operation a.
One may similarly obtain
Z Z
a d3 r = n a d2 r Cross-Product Theorem (1.34)

which is not Stokes Theorem even if it involves the curl operation a.


1.5. DIVERGENCE OF VECTOR (AND TENSOR) FIELDS 9

Stokes Theorem: For some finite open (possibly curved) surface S having normal n at each point
on S and bounded by a closed contour

Z I
n ( a) dS = a dl Stokes Theorem (1.35)
S

where dl is the
H infinitesimal length vector tangent to points on . Note that the
integral sign means that we start the integral on the closed contour at one
pont and finish the integral at that same point. We will use Stokes Theorem in
interpreting the physical meaning of the curl differential operator.

1.5 Divergence of Vector (and Tensor) Fields


The divergence operation is the dot product between the gradient operator and either a vector or tensor
 

a= x + y + z [ax x
+ ay y + az z]
. (1.36)
x y z
In carrying out the products and derivatives, note that in Cartesian coordinates, the base vectors are uniform
constants and have the property that x j /xi = 0. Note that this is not true for other base vectors in
curvilinear coordinates (e.g., cylindrical and spherical). Since the derivatives of the base vectors are zero in
Cartesian coordinates we have
ax ay az ai
a= + + = (1.37)
x y z xi
with summation of the index i as always.
Our goal now is to obtain the physical meaning of a at a point r. To do so, construct a tiny volume V
around the point r that is bounded by a surface S and use the Divergence Theorem:
Z Z
1 1
a dV = n a dS. (1.38)
V V V S
For V sufficiently small, a constant in V, so
Z Z
1 1
a = lim a dV = lim n a dS =
b accumulation of the physical quantity carried by a.
V 0 V V V 0 V S

Imagine the vector a as a flux of some type carrying a physical quantity with it. If the flux into a small
element V is greater than the flux out of V , which is what the surface integral allows for, then the divergence
of this flux is non-zero. The divergence of a vector field is associated with the idea of accumulation (if the
divergence is negative) or depletion (if the divergence is positive) in a tiny region surrounding the point in
question.
Thus to see a, we imagine what is happening on S:

a 6= 0 a=0
<0
So if the vector field a is larger on one side of the small element V compared to the other. the divergence is
non-zero. If the divergence is positive it means n a (the flux of vector a across ) was larger on one side
of the volume compared to the other and that the physical quantity being carried by a is being depleted. If
the divergence is negative, the physical quantity is accumulating.
10 CHAPTER 1. SOME ACTORS AND ACTS

There are two ways we commonly use to visualize a vector field:

little vectors at each point current lines

For current lines, the direction of the field at a point is tangent to the current line at that point. The
amplitude of the field is given by the density of current lines in the neighborhood surrounding the point.
Wherever a line starts, a 6= 0 as depicted below.

a=0 a 6= 0

In particular, imagine the electric field around a point charge q:

E =0

E 6= 0

So no lines start and stop in each element and the divergence is zero except when the element includes the
charge itself where all field lines begin and the divergence is non-zero.

What about the divergence of a tensor? We procede like we did for a vector
T = a vector given by

  Txx Txy Txz

= , , Tyx Tyy Tyz
x y z
Tzx Tzy Tzz
h T
xx Tyx Tzx Txy Tyy Tzy Txz Tyz Tzz iT
= + + , + + , + + (1.39)
x y z x y z x y z
| {z } | {z } | {z }
x component y component z component
 

= x i (Tjk xj x
k ) (summation over repeated indices)
xi
Tjk
= (xi x
j ) x
k but x i x
j = 0 unless j = i so
xi
Tik
= xk = T .
xi
1.5. DIVERGENCE OF VECTOR (AND TENSOR) FIELDS 11

To visualize the meaning of the divergence of a tensor, we can use our same device of considering a tiny
element V surrounding the point where the vector T is being evaluated and obtain (because V is so small
that T is uniform inside of V )
Z Z
1 1
T = lim T dV = lim n T dS. (1.40)
V 0 V V V 0 V S
Think of the vector n T as being a force so that if the force acting on one side of the element is larger than
the force acting on the other side (which is what the surface integral allows for), there is a net force acting
on the element characterized by T 6= 0.

SIDE BOX 1: How do you distribute the gradient operator in an expression involving several vec-
torial or tensorial terms? What is the equivalent of the chain rule ()/x = /x + /x?
As an example, lets consider the specific expression

(ab T ) (1.41)

where a and b are both vectors and T is a 2nd-order tensor. To distribute the derivative in this ex-
pression, we work in Cartesian coordinates in which the base vectors x i are uniform constants (do not
change based on where you are situated in space). In curvilinear coordinates, some of the base vectors
change their orientation with position in space (imagine the radial vector in spherical coordinates)
and one must take derivatives of both the components and the base vectors. It greatly simplifies our
analysis to work in Cartesians where x j /xi = 0 for all base vectors and coordinate directions.
For each vectorial or tensorial term in the expression, we use a different set of Cartesian-coordinate
indices to write out

(ab T ) = x
i (aj x k Tlm x
j bk x lx
m) (1.42)
xi

= i x
(aj bk Tlm ) (x k x
j ) (x l) x
m. (1.43)
xi
In passing from the first line to the second, we pulled out the components while preserving the posi-
tion of the base vectors and scalar products between the base vectors. Next, because of the nature of
the scalar product in Cartesian coordinates, we have that j = i and l = k. This allows us to write

(ab T ) = m
(ai bk Tkm ) x (1.44)
xi
    
ai bk Tkm
= bk Tkm + ai Tkm + ai bk m.
x (1.45)
xi xi xi

In going from the first to second expression, we just employed the usual derivative chain rule.
The final step is what students often find the most difficult. One must look at Eq. (1.45) and identify
the equivalent expression in bold face. So you have to make identifications like ai /xi = a and
m = b T . Carrying this out, we obtain at last
bk Tkm x

(ab T ) = ( a) b T + a (b) T + b (a T ) . (1.46)

Once we type the final expression in bold face (or write by hand the expression with squiggly un-
derscores), it applies to any curvilinear coordinates and is a generally valid identity not limited to
Cartesian coordinates. Note that T is an example of a third-order tensor. We have thus determined
how to distribute the operator onto the vectors and tensors in a multi-term expression involving
scalar products. You will practice proving such tensor calculus identities in some homework problems.
12 CHAPTER 1. SOME ACTORS AND ACTS

SIDE BOX 2: What does a look like when written out in cylindrical coordinates?

In cylindricals, we have that the gradient operator goes as


1
= r + + z . (1.47)
r r z

A small distance in the direction goes as rd and that is why there is an r in the derivative term.
Further, the r and base vectors change their orientation with changing . As such, we also have
(without pausing to demonstrate)
r
= and = r. (1.48)

Thus, when we calculate
  h
1 i
a = r + + z ar r + a + az z (1.49)
r r z

and carry out all the derivatives including the derivatives of the base vectors and then perform the dot
products, we end up with
 
ar 1 a az
a= + ar + + (1.50)
r r a z
1 1 a az
= (rar ) + + . (1.51)
r r r z
So performing operations between the operator and vectors and tensors in curvilinear coordinates
is more complicated than in Cartesians not just because the operator itself is a bit more complicated
but because the base vectors have non-zero derivatives that contribute.

1.6 Rotation or Curl of Vector Fields


The curl is the vector product between the gradient operator and a vector given by

y z
x      
az ay az ax ay ax
a = x y
z = x y + z (1.52)
ax y z x z x y
ay az

The curl is also sometimes called the rotation.


For the physical meaning of the curl operation, we use Stokes Theorem in the limit that the open surface S
at a point is so small that a can be taken as a constant over S
Z I
1 1
n a = lim n ( a) dS = lim a dl. (1.53)
S0 S S S0 S

Thus, if the vector field a is tangent in places to the curve and if this tangential component is larger on
one side of S compared to the other so that integral over is non-zer, then we will have a 6= 0. We can
associate this with the idea that the field has a rotation in the vicinity of S.
1.6. ROTATION OR CURL OF VECTOR FIELDS 13

To visualize the curl, we use a water wheel as depicted to the left. We immerse
the wheel in our field a that we imagine to be the flow of water regardless of what
a really corresponds to. We change the orientation of the axle of the wheel and
observe how fast the wheel is moving if at all. The direction n of the axle at which
the wheel turns the fastest gives the direction of a, while the rate of rotation
gives | a|. We use the right-hand rule to make sure we have the sign of n
(which is the direction of a) correct.

So to mentally investigate the curl (or rotation) associated with some field, we imagine the vector field to be
a flow field and probe the field with our water wheel to see in what orientation the wheel turns the fastest
if, indeed, it can turn at all:

(a) Uniform Field a=0


this field cannot
make the wheel
turn.

a 6= 0
(b)
a = | a| y

Note that a water wheel is not the same as a propellar. A propellar would turn even in case (a) above of a
uniform field.

CONCLUSIONS: To conclude these discussions of how to understand and visualize non-zero values of
the divergence and curl of a vector field, we can define longitudinal and transverse spatial variations of
a vector field. Longitudinal variations are those in the same direction of the vector field and transverse
variations are those in a direction perpendicular to the vector field. Longitudinal variations (the direc-
a in the direction of the vector field) lead to a non-zero divergence but do not, by
tional derivative a
themselves, contribute to the curl. Similarly, tranverse variations lead to a non-zero curl but do not,
by themselves, contribute to the divergence. The decomposition of a vector field into longitudinal and
transverse variations will correspond to P-waves and S-waves respectively when we discuss seismology.

We can also perform the curl differential operation using the Levi-Civita tensor in the following way:


a = 3  : a = ijk x
ix
j x
k : x
l m
am x
xl

aj
= eijk i
x (1.54)
xk
     
a3 a2 a1 a3 a2 a1
1
=x 2
+x 3
+x
x2 x3 x3 x1 x1 x2

Being able to do curl operations using the dot product makes it possible to prove many useful things about
curl operations.
14 CHAPTER 1. SOME ACTORS AND ACTS

In continuum mechanics, we will make frequent use of the following identity:

u = ( u) 2 u (1.55)

We can prove this identity as follows:


 

u = ijk x
ix
j x
k : l mno x
x mx
nx
o : x
p q
uq x (1.56)
xl xp
2 uq
= ijk mno x k x
i (x j x
l ) (x o x
m ) (x n x
p ) (x q )
xl xp
So l = k, m = j, p = o, and q = n, to give
2 un
u = ijk jno i
x
xk xo
2 uk 2 ui
 
= ijk jki + ijk jik 2 x i sum over k and j (1.57)
xi xk xk
2
 
uk
=xi i
ui x
xi xk x2k
= ( u) 2 u

Using 3 , it is also straightforward to show = 0 and ( a) = 0 which are highly useful relations
that we will use, for example, when we manipulate Maxwells electromagnetic equations.

1.7 The Dirac Delta Function


The Dirac delta is used to represent fields that are highly concentrated in space (or time):
(
if x = x1
(x x1 ) = (1.58)
0 if x 6= x1

in such a way that if a < x1 < b, then


Z b
(x x1 ) dx = 1 (no units) (1.59)
a

In some sense, we have that when x = x1 , then (0) = dx1 (i.e., the Dirac delta function spikes up to a
large value when its argument is zero). In this case one has

Z b
f (x) (x x1 ) dx = f (x1 ) (1.60)
a

which is called the sifting property. In the integration process, the Dirac delta function samples f where the
argument of the Dirac goes to zero. Note as well that from Eq. (1.59) [or equivalently the sifting property],
we necessarily have that (x) has physical of units of x1 whatever the physical units of x are. This is
important to remember in physics applications.
If Eq. (1.58) does not seem like the definition of a well-behaved function, we can alternatively define (xx1 )
as the limit of better-defined smooth functions:
n 1
(x x1 ) = lim
n 1 + n (x x1 )2
2
n 2 2
= lim n (xx1 ) (1.61)
n
1.7. THE DIRAC DELTA FUNCTION 15

and so forth, which satisfy the integral constraint of Eq. (1.59) in the limit of large n. In this way, one can
take deriviatives of (x x1 ):
d (x x1 ) n3 2 2
= lim 2(x x1 ) n (xx1 ) (1.62)
dx n
We will use the Dirac to represent highly concentrated fields in nature; however, in each application to a
physics problem, the Dirac delta function will be integrated over which due to the defining integral of Eq.
(1.59) leads to well-behaved results no matter how concentrated one makes the Dirac when its argument is
zero.
Visually, one can see the effect of taking the derivative of a Dirac delta by using Eqs. (1.61) and (1.62):

d (xx1 )
(x x1 ) dx

Also, we can define the step function S(x x1 ) as


Z x (
1 x x1
S(x x1 ) = (x0 x1 ) dx0 = . (1.63)
0 x < x1
Thus, we also have

d
(x x1 ) = S(x x1 ) (1.64)
dx

Note that if a < x1 < b,


Z b Z b h 
d (x x1 ) d i d f (x)
f (x) dx = f (x) (x x1 ) (x x1 ) dx (1.65)
a dx a dx dx

d f (x)

=
dx x=x1
which can be called the derivative-sifting property. This result generalizes to an arbitrary number of deriva-
tives n as Z b
dn (x x1 ) n

n d f (x)

f (x) n
dx = (1) n
. (1.66)
a dx dx x=x1

IMPORTANT: For a field concentrated at a point r1 in 3D space, we will make frequent use of the
notation
(r r1 ) =
b (x x1 ) (y y1 ) (z z1 ). (1.67)
Note that (r r1 ) has units of inverse length cubed, because () with some scalar, has units
of 1 as Eq. (1.59) makes clear.

You might also like