Math 107 Fall 2011 Notes
Math 107 Fall 2011 Notes
Lecture Notes
Adolfo J. Rumbos
⃝
c Draft date November 23, 2011
2
Contents
2 Euclidean Space 7
2.1 Definition of 𝑛–Dimensional Euclidean Space . . . . . . . . . . . 7
2.2 Spans, Lines and Planes . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Dot Product and Euclidean Norm . . . . . . . . . . . . . . . . . 11
2.4 Orthogonality and Projections . . . . . . . . . . . . . . . . . . . 13
2.5 The Cross Product in ℝ3 . . . . . . . . . . . . . . . . . . . . . . 18
2.5.1 Defining the Cross–Product . . . . . . . . . . . . . . . . . 20
2.5.2 Triple Scalar Product . . . . . . . . . . . . . . . . . . . . 24
3 Functions 27
3.1 Types of Functions in Euclidean Space . . . . . . . . . . . . . . . 27
3.2 Open Subsets of Euclidean Space . . . . . . . . . . . . . . . . . . 28
3.3 Continuous Functions . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.1 Images and Pre–Images . . . . . . . . . . . . . . . . . . . 35
3.3.2 An alternate definition of continuity . . . . . . . . . . . . 35
3.3.3 Compositions of Continuous Functions . . . . . . . . . . . 37
3.3.4 Limits and Continuity . . . . . . . . . . . . . . . . . . . . 38
4 Differentiability 41
4.1 Definition of Differentiability . . . . . . . . . . . . . . . . . . . . 42
4.2 The Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3 Example: Differentiable Scalar Fields . . . . . . . . . . . . . . . . 44
4.4 Example: Differentiable Paths . . . . . . . . . . . . . . . . . . . . 49
4.5 Sufficient Condition for Differentiability . . . . . . . . . . . . . . 51
4.5.1 Differentiability of Paths . . . . . . . . . . . . . . . . . . . 51
4.5.2 Differentiability of Scalar Fields . . . . . . . . . . . . . . . 53
4.5.3 𝐶 1 Maps and Differentiability . . . . . . . . . . . . . . . . 55
4.6 Derivatives of Compositions . . . . . . . . . . . . . . . . . . . . . 56
5 Integration 61
5.1 Path Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.1.1 Arc Length . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3
4 CONTENTS
B Reparametrizations 107
Chapter 1
Then ∫ 𝑏
𝑓 (𝑥)d𝑥 = 𝐹 (𝑏) − 𝐹 (𝑎). (1.1)
𝑎
The main goal of this course is to extend this result to higher dimensions. In
order to indicate how we intend to do so, we first re-write the integral in (1.1)
as follows:
First denote the interval [𝑎, 𝑏] by 𝑀 ; then, its boundary, denoted by ∂𝑀 , consists
of the end–points 𝑎 and 𝑏 of the interval; thus,
∂𝑀 = {𝑎, 𝑏}.
1 Recall that a function 𝑓 : 𝐼 → ℝ is continuous at 𝑐 ∈ 𝐼, if (i) 𝑓 (𝑐) is defined, (ii) lim 𝑓 (𝑥)
𝑥→𝑐
exists, and (iii) lim 𝑓 (𝑥) = 𝑓 (𝑐).
𝑥→𝑐
2 Recall 𝑓 (𝑥) − 𝑓 (𝑐)
that a function 𝑓 : 𝐼 → ℝ is differentiable at 𝑐 ∈ 𝐼, if lim exists.
𝑥→𝑐 𝑥−𝑐
5
6 CHAPTER 1. MOTIVATION FOR THE COURSE
The reason for doing this change in notation is so that later on we can talk
about integrals over regions 𝑀 in Euclidean space, and not just integrals over
intervals. Thus, the concept of the integral will also have to be expanded. To
see how this might come about, we discuss briefly how the right–hand side the
expression in (1.1) might also be expressed as an integral.
Rewrite the right–hand side of (1.1) as the sum
thus, we are adding the values of the function 𝐹 on the boundary of 𝑀 taking
into account the convention that, as we do the integration on the left–hand side
of (1.1), we go from left to right along the interval [𝑎, 𝑏]; hence, as we integrate,
“we leave 𝑎” (this explains the −1 in front of 𝐹 (𝑎)) and “we enter 𝑏” (hence the
+1 in from of 𝐹 (𝑏)). Since integration of a function is, in some sense, the sum
of its values over a certain region, we are therefore led to suggesting that the
right–hand side in (1.1) may be written as:
∫
𝐹.
∂𝑀
Thus the result of the Fundamental Theorem of Calculus in equation (1.1) may
now be written in a more general form as
∫ ∫
𝑑𝐹 = 𝐹. (1.2)
𝑀 ∂𝑀
This is known as the Generalized Stokes’ Theorem and a precise state of this
theorem will be given later in the course. It says that under certain conditions
on the sets 𝑀 and ∂𝑀 , and the “integrands,” also to be made precise later
in this course, integrating the “differential” of “something” over some “set,” is
the same as integrating that “something” over the boundary of the set. Before
we get to the stage at which we can state and prove this generalized form of
the Fundamental Theorem of Calculus, we will need to introduce concepts and
theory that will make the terms “something,” “set” and “integration on sets”
make sense. This will motivate the topics that we will discuss in this course.
Here is a broad outline of what we will be studying.
Euclidean Space
7
8 CHAPTER 2. EUCLIDEAN SPACE
∙ Vector Addition
⎛ ⎞ ⎛ ⎞
𝑥1 𝑦1
⎜ 𝑥2 ⎟ ⎜ 𝑦2 ⎟
Given 𝑣 = ⎜ . ⎟ and 𝑤 = ⎜ . ⎟ , the vector sum 𝑣 + 𝑤 or 𝑣 and 𝑤 is
⎜ ⎟ ⎜ ⎟
⎝ .. ⎠ ⎝ .. ⎠
𝑥𝑛 𝑦𝑛
⎛ ⎞
𝑥1 + 𝑦1
⎜ 𝑥2 + 𝑦2 ⎟
𝑣+𝑤 =⎜
⎜ ⎟
.. ⎟
⎝ . ⎠
𝑥𝑛 + 𝑦𝑛
∙ Scalar Multiplication
⎛ ⎞
𝑥1
⎜ 𝑥2 ⎟
Given a real number 𝑡, also called a scalar, and a vector 𝑣 = ⎜ . ⎟ the
⎜ ⎟
⎝ .. ⎠
𝑥𝑛
scaling of 𝑣 by 𝑡, denoted by 𝑡𝑣, is given by
⎛ ⎞
𝑡𝑥1
⎜ 𝑡𝑥2 ⎟
𝑡𝑣 = ⎜ . ⎟
⎜ ⎟
⎝ .. ⎠
𝑡𝑥𝑛
Remark 2.1.2. In some texts, vectors are denoted with an arrow over the
symbol for the vector; for instance, →−𝑣, →
−𝑟 , etc. In the text that we are using
this semester, vectors are denoted in bold face type, v, r, etc. For the most part,
we will do away with arrows over symbols and bold face type in these notes,
lectures, and homework assignments. The context will make clear whether a
given symbol represents a point, a number, a vector, or a matrix.
Geometrically, if 𝑣 is not the zero vector in ℝ𝑛 , span{𝑣} is the line through the
origin on ℝ𝑛 in the direction of the vector 𝑣.
If 𝑃 is a point in ℝ𝑛 and 𝑣 is a non–zero vector also in ℝ𝑛 , then the line
through 𝑃 in the direction of 𝑣 is the set
−−→ −−→
𝑂𝑃 + span{𝑣} = {𝑂𝑃 + 𝑡𝑣 ∣ 𝑡 ∈ ℝ}.
2.2. SPANS, LINES AND PLANES 9
⎞ ⎛
2
Example 2.2.1 (Parametric Equations of a line in ℝ3 ). Let 𝑣 = ⎝−3⎠ be a
1
vector in ℝ3 and 𝑃 the point with coordinates (1, 0 − 1). Find the line through
𝑃 in the direction of 𝑣.
Solution: The line through 𝑃 in the direction of 𝑣 is the set
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎫
⎨ 𝑥 𝑥 1 2 ⎬
⎝𝑦 ⎠ ∈ ℝ3 ⎝𝑦 ⎠ = ⎝ 0 ⎠ + 𝑡 ⎝−3⎠ , 𝑡 ∈ ℝ
𝑧 𝑧 −1 1
⎩ ⎭
or ⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎫
⎨ 𝑥 𝑥 1 + 2𝑡 ⎬
⎝𝑦 ⎠ ∈ ℝ3 ⎝𝑦 ⎠ = ⎝ −3𝑡 ⎠ , 𝑡 ∈ ℝ
𝑧 𝑧 −1 + 𝑡
⎩ ⎭
⎛ ⎞
𝑥
Thus, for a point ⎝𝑦 ⎠ to be on the line, 𝑥, 𝑦 and 𝑧 must satisfy
𝑧
the equations ⎧
⎨ 𝑥 = 1 + 2𝑡;
𝑦 = −3𝑡;
𝑧 = −1 + 𝑡,
⎩
In some cases we are interested in the directed line segment from a point
𝑃1 (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ) to a point 𝑃1 (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ) in ℝ𝑛 . We will denote this set
by [𝑃1 𝑃2 ]; so that,
−−→ −−−→
[𝑃1 𝑃2 ] = {𝑂𝑃1 + 𝑡𝑃1 𝑃2 ∣ 0 ⩽ 𝑡 ⩽ 1}.
is a plane through the origin containing the points located by the vectors 𝑣1 and
𝑣2 .
If 𝑃 is a point in ℝ3 , the plane through 𝑃 generated by the linearly inde-
pendent vectors 𝑣1 and 𝑣2 , also in ℝ3 , is given by
−−→ −−→
𝑂𝑃 + span{𝑣1 , 𝑣2 } = {𝑂𝑃 + 𝑡𝑣1 + 𝑠𝑣2 ∣ 𝑡, 𝑠 ∈ ℝ}.
⎛ ⎞ ⎛ ⎞
2 6
Example 2.2.2 (Equations of planes ℝ3 ). Let 𝑣1 = ⎝−3⎠ and 𝑣2 = ⎝ 2 ⎠
1 −3
3
be vectors in ℝ and 𝑃 the point with coordinates (1, 0 − 1). Give the equation
of the plane through 𝑃 spanned by the vectors 𝑣1 and 𝑣2 .
for 𝑡 and 𝑠.
Using Gaussian elimination, we get can determine conditions on 𝑥,
𝑦 and 𝑧 that will allows us to solve for 𝑡 and 𝑠:
3 ∣ 𝑥−1
⎛ ⎞ ⎛ ⎞
2 6 ∣ 𝑥−1 1 2
⎝−3 2 ∣ 𝑦 ⎠ → ⎝−3 2 ∣ 𝑦 ⎠→
1 −3 ∣ 𝑧 + 1 1 −3 ∣ 𝑧 + 1
𝑥−1 𝑥−1
⎛ ⎞ ⎛ ⎞
1 3 ∣ 2 1 3 ∣ 2
3 3 1
⎝ 0 11 ∣ 2 (𝑥 − 1) + 𝑦 ⎠→⎝ 0 1 ∣ 22 (𝑥 − 1) + 11 𝑦
⎠→
0 −6 ∣ − 12 (𝑥 − 1) + (𝑧 + 1) 0 −1 ∣ 1 1
− 12 (𝑥 − 1) + 6 (𝑧 + 1)
𝑥−1
⎛ ⎞
1 3 ∣ 2
3 1
⎝ 0 1 ∣ 22 (𝑥− 1) + 11 𝑦 ⎠
7 1 1
0 0 ∣ 132 (𝑥 − 1) + 11 𝑦 + 6 (𝑧 + 1)
2.3. DOT PRODUCT AND EUCLIDEAN NORM 11
Thus, for the system to be solvable for 𝑡 and 𝑠, the third row must
be a row of zeros. We therefore get the equation
7 1 1
(𝑥 − 1) + 𝑦 + (𝑧 + 1) = 0
132 11 6
or
7(𝑥 − 1) + 12(𝑦 − 0) + 22(𝑧 + 1) = 0.
This is the equation of the plane. □
The superscript 𝑇 in the above definition indicates that the column vector
𝑣 has been transposed into a row vector.
The inner or dot product defined above satisfies the following properties
which can be easily checked:
(i) Symmetry: 𝑣 ⋅ 𝑤 = 𝑤 ⋅ 𝑣
𝑧
6
𝑃
@
H@ 𝑤
:
r
HH
@H
@ @HH
@R HH
@
𝑣@@ 𝑡𝑣 HH
𝑥
@
@
R
@ j 𝑦
H
@
@ span{𝑣}
that is, 𝑓 (𝑡) is the square of the distance from 𝑃 to any point on the line through
𝑂 in the direction of 𝑣. We wish to minimize this function.
Observe that 𝑓 (𝑡) can be written in terms of the dot product as
𝑓 (𝑡) = (𝑤 − 𝑡𝑣) ⋅ (𝑤 − 𝑡𝑣),
which can be expanded by virtue of the properties of the inner product and the
definition of the Euclidean norm into
𝑓 (𝑡) = ∥𝑤∥2 − 2𝑡𝑣 ⋅ 𝑤 + 𝑡2 ∥𝑣∥2 .
Thus, 𝑓 (𝑡) is a quadratic polynomial in 𝑡 which can be shown to have an absolute
minimum when
𝑣⋅𝑤
𝑡= .
∥𝑣∥2
Thus, the point on span{𝑣} which is closest to 𝑃 is the point
𝑣⋅𝑤
𝑣,
∥𝑣∥2
−−→
where 𝑤 = 𝑂𝑃 .
The distance form 𝑃 to the line (i.e., the shortest distance) is then
𝑣⋅𝑤
𝑣−𝑤 .
∥𝑣∥2
Remark 2.4.2. The argument of the previous example can be used to show that
the point on the line
−−→
𝑂𝑃𝑜 + span{𝑣},
for a given point 𝑃𝑜 , which is the closest to 𝑃 is given by
−−→ 𝑣 ⋅ 𝑤
𝑂𝑃𝑜 + 𝑣,
∥𝑣∥2
−−→
where 𝑤 = 𝑃𝑜 𝑃 , and the distance from 𝑃 to the line is
−−→ 𝑣 ⋅ 𝑤
𝑂𝑃𝑜 + 𝑣−𝑤 .
∥𝑣∥2
Definition 2.4.3 (Orthogonality). Two vectors 𝑣 and 𝑤 in ℝ𝑛 are said to be
orthogonal, or perpendicular, if
𝑣 ⋅ 𝑤 = 0.
Definition 2.4.4 (Orthogonal Projection). The vector
𝑣⋅𝑤
𝑣
∥𝑣∥2
is called the orthogonal projection of 𝑤 onto 𝑣. We denote it by 𝑃𝑣 (𝑤). Thus,
(𝑣 ⋅ 𝑤)
𝑃𝑣 (𝑤) = 𝑣.
∥𝑣∥2
2.4. ORTHOGONALITY AND PROJECTIONS 15
−−→
𝑃𝑣 (𝑤) is called the orthogonal projection of 𝑤 = 𝑂𝑃 onto 𝑣 because it lies
along a line through 𝑃 which is perpendicular to the direction of 𝑣. To see why
this is the case compute
𝑤 = 𝑃𝑣 (𝑤) + (𝑤 − 𝑃𝑣 (𝑤));
that is, the sum of a vector parallel to 𝑣 and another vector perpendicular to 𝑣.
This is known as the orthogonal decomposition of 𝑤 with respect to 𝑣.
Example 2.4.5. Let 𝐿 denote the line given parametrically by the equations
⎧
⎨ 𝑥 = 1−𝑡
𝑦 = 2𝑡 (2.1)
𝑧 = 2 + 𝑡,
⎩
for 𝑡 ∈ ℝ. Find the point on the line, 𝐿, which is closest to the point 𝑃 (1, 2, 0)
and compute the distance from 𝑃 to 𝐿.
(𝑣 ⋅ 𝑤) 2 1
𝑃𝑣 (𝑤) = 𝑣 = 𝑣 = 𝑣.
∥𝑣∥2 6 3
16 CHAPTER 2. EUCLIDEAN SPACE
so that ⎛ ⎞
1/3
dist(𝑃, 𝐿) = ⎝ 4/3 ⎠
−7/3
⎛ ⎞
1
1 ⎝ ⎠
= 4
3
−7
√
1√ 66
= 1 + 16 + 49 = .
3 3
□
1 1
∥ˆ
𝑣∥ = 𝑣 = ∥𝑣∥ = 1.
∥𝑣∥ ∥𝑣∥
Let 𝑃 denote a point which is not on the plane 𝐻. Find the shortest distance
from the point 𝑃 to 𝐻.
Solution: Let 𝑃𝑜 (𝑥𝑜 , 𝑦𝑜 , 𝑧𝑜 ) be any point in the plane, 𝐻, and define
−−→
the vector, 𝑤 = 𝑃𝑜 𝑃 , which goes from the point 𝑃𝑜 to the point 𝑃 .
The shortest distance from 𝑃 to the plane will be the norm of the
projection of 𝑤 onto the orthogonal direction vector,
⎛ ⎞
𝑎
𝑛 = ⎝ 𝑏⎠ ,
𝑐
18 CHAPTER 2. EUCLIDEAN SPACE
2𝑥 + 3𝑦 + 6𝑧 = 6.
∣𝑤 ⋅ 𝑛∣
dist(𝑃, 𝐻) = ,
∥𝑛∥
where ⎛ ⎞
2
𝑛 = ⎝ 3⎠ ,
6
so that
𝑤 ⋅ 𝑛 = 12,
and √
∥𝑛∥ = 4 + 9 + 36 = 7.
Consequently,
12
dist(𝑃, 𝐻) = .
7
□
𝑦
𝑏2 𝑤
A
A
Aℎ
A
A
𝑎2 A
𝑣
*
𝑏1 𝑎1 𝑥
Figure 2.5.2 shows shows a sketch of the arrows representing 𝑣 and 𝑤 for the
special case in which they lie in the first quadrant of the 𝑥𝑦–plane.
We would like to compute the area of the parallelogram, 𝑃 (𝑣, 𝑤), determined
by 𝑣 and 𝑤. This may be computed as follows:
where ℎ may be obtained as ∥𝑤 − 𝑃𝑣 (𝑤)∥; that is, the distance from 𝑤 to its
orthogonal projection along 𝑣. Squaring both sides of the previous equation we
have that
(𝑣 ⋅ 𝑤)2
( )
(𝑣 ⋅ 𝑤)
= ∥𝑣∥2 ∥𝑤∥2 − 2𝑤 ⋅ 𝑣 +
∥𝑣∥2 ∥𝑣∥2
(𝑣 ⋅ 𝑤)2
( )
(𝑣 ⋅ 𝑤)
= ∥𝑣∥2 ∥𝑤∥2 − 2 𝑤 ⋅ 𝑣 +
∥𝑣∥2 ∥𝑣∥2
(𝑣 ⋅ 𝑤)2 (𝑣 ⋅ 𝑤)2
( )
= ∥𝑣∥2 ∥𝑤∥2 − 2 +
∥𝑣∥2 ∥𝑣∥2
= 𝑎21 𝑏21 + 𝑎21 𝑏22 + 𝑎22 𝑏21 + 𝑎22 𝑏22 − (𝑎21 𝑏21 + 2𝑎1 𝑏1 𝑎2 𝑏2 + 𝑎22 𝑏22 )
= (𝑎1 𝑏2 − 𝑎2 𝑏1 )2 .
(2.4)
Taking square roots on both sides, we get
Observe that the expression in the absolute value on the right-hand side of the
previous equation is the determinant of the matrix:
( )
𝑎1 𝑏1
.
𝑎2 𝑏2
Observe that this formula works even in the case in which 𝑣 and 𝑤 are not
linearly independent. In this case we get that the area of the parallelogram
determined by the two vectors is 0.
𝑦
𝑏2 𝑤
A
A
Aℎ
A
A
𝑎2 A
𝑣
*
𝑏1 𝑎1 𝑥
𝑎1 𝑏1 ˆ
𝑣×𝑤 = 𝑘.
𝑎2 𝑏2
Observe that, since the determinant of the transpose of a matrix is the same
as that of the matrix, we can also write
𝑎1 𝑎2 ˆ
𝑣×𝑤 = 𝑘, (2.5)
𝑏1 𝑏2
for vectors ⎛ ⎞ ⎛ ⎞
𝑎1 𝑏1
𝑣 = ⎝𝑎2 ⎠ and 𝑤 = ⎝𝑏2 ⎠
0 0
lying in the 𝑥𝑦–plane.
In general, the cross product of the vectors
⎛ ⎞ ⎛ ⎞
𝑎1 𝑏1
𝑣 = ⎝𝑎2 ⎠ and 𝑤 = ⎝𝑏2 ⎠
𝑎3 𝑏3
in ℝ3 is the vector
𝑎2 𝑎3 ˆ 𝑎1 𝑎3 ˆ 𝑎 𝑎2 ˆ
𝑣×𝑤 = 𝑖− 𝑗+ 1 𝑘, (2.6)
𝑏2 𝑏3 𝑏1 𝑏3 𝑏1 𝑏2
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
where ˆ𝑖 = ⎝0⎠ , ˆ 𝑗 = ⎝1⎠ , and ˆ 𝑘 = ⎝0⎠ are the standard basis vectors in
0 0 1
ℝ3 .
Observe that if 𝑎3 = 𝑏3 = 0 in definition on 𝑣 × 𝑤 in (2.6), we recover the
expression in (2.5),
𝑎 𝑎2 ˆ
𝑣×𝑤 = 1 𝑘
𝑏1 𝑏2
for the cross product of vectors lying entirely in the 𝑥𝑦–plane.
2.5. THE CROSS PRODUCT IN ℝ3 23
In the remainder of this section, we verify that the cross product of two
vectors, 𝑣 and 𝑤, in ℝ3 defined in the (2.6) does indeed satisfies the properties
listed at the beginning of the section. To check that 𝑣 × 𝑤 is orthogonal to the
plane spanned by 𝑣 and 𝑤, write
⎛ ⎞ ⎛ ⎞
𝑎1 𝑏1
𝑣 = ⎝𝑎2 ⎠ and 𝑤 = ⎝𝑏2 ⎠
𝑎3 𝑏3
𝑣 ⋅ (𝑣 × 𝑤) = 𝑣 𝑇 (𝑣 × 𝑤)
⎛ ⎞
𝑎2 𝑎3
⎜
⎜ 𝑏2 𝑏3 ⎟
⎟
⎜ ⎟
⎜ ⎟
( ) ⎜ 𝑎1 𝑎3 ⎟
= 𝑎1 𝑎2 𝑎3 ⎜⎜− 𝑏1
⎟
⎜ 𝑏3 ⎟
⎟
⎜ ⎟
⎜ ⎟
⎝ 𝑎1 𝑎2 ⎠
𝑏1 𝑏2
so that
𝑎2 𝑎3 𝑎 𝑎3 𝑎 𝑎2
𝑣 ⋅ (𝑣 × 𝑤) = 𝑎1 − 𝑎2 1 + 𝑎3 1 .
𝑏2 𝑏3 𝑏1 𝑏3 𝑏1 𝑏2
We recognize in the right–hand side of equation (??) the expansion along the
first row of the determinant
𝑎1 𝑎2 𝑎3
𝑎1 𝑎2 𝑎3 ,
𝑏1 𝑏2 𝑏3
which is 0 since the first two rows are the same. Thus,
𝑣 ⋅ (𝑣 × 𝑤) = 0,
𝑏1 𝑏2 𝑏3
𝑤 ⋅ (𝑣 × 𝑤) = 𝑎1 𝑎2 𝑎3 = 0,
𝑏1 𝑏2 𝑏3
The calculations displayed in (2.4) then show that (2.7) can we written as
𝑛=𝑣×𝑤
6
𝑢
ℎ
𝑤
*
-
𝑣
First, observe that the volume of the parallelepiped, 𝑃 (𝑣, 𝑤, 𝑢), drawn in
Figure 2.5.4 is the area of the parallelogram spanned by 𝑣 and 𝑤 times the
height, ℎ, of the parallelepiped:
𝑢⋅𝑛
ℎ = ∥𝑃𝑛 (𝑢)∥ = 𝑛 ,
∥𝑛∥2
where
𝑛 = 𝑣 × 𝑤.
2.5. THE CROSS PRODUCT IN ℝ3 25
The scalar, 𝑢⋅(𝑣 ×𝑤), in the right–hand side of the equation in (2.9) is called
the triple scalar product of 𝑢, 𝑣 and 𝑤.
Given three vectors
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
𝑐1 𝑎1 𝑏1
𝑢 = ⎝𝑐2 ⎠ , 𝑣 = ⎝𝑎2 ⎠ and 𝑤 = ⎝𝑏2 ⎠
𝑐3 𝑎3 𝑏3
𝑎2 𝑎3 𝑎 𝑎3 𝑎 𝑎2
𝑢 ⋅ (𝑣 × 𝑤) = 𝑐1 − 𝑐2 1 + 𝑐3 1 ,
𝑏2 𝑏3 𝑏1 𝑏3 𝑏1 𝑏2
or
𝑐1 𝑐2 𝑐3
𝑢 ⋅ (𝑣 × 𝑤) = 𝑎1 𝑎2 𝑎3 ;
𝑏1 𝑏2 𝑏3
that is, 𝑢 ⋅ (𝑣 × 𝑤) is the the determinant of the 3 × 3 matrix whose rows are the
vector 𝑢, 𝑣 and 𝑤, in that order. Since the determinant of the transpose of a
matrix is the same as the determinant of the original matrix, we may also write
𝑢 ⋅ (𝑣 × 𝑤) = det[ 𝑢 𝑣 𝑤 ],
the determinant of the 3 × 3 matrix whose columns are the vector 𝑢, 𝑣 and 𝑤,
in that order.
26 CHAPTER 2. EUCLIDEAN SPACE
Chapter 3
Functions
𝐹 : 𝐷 → ℝ𝑚
and call 𝐷 the domain of 𝐹 ; that is, the set where the function is defined.
1
𝑓 (𝑥, 𝑦) = √
1 − 𝑥2 − 𝑦 2
There are different types of functions that we will be studying in this course.
Some of the types have received traditional names, and we present them here.
𝐹 : 𝐷 → ℝ𝑛
is called a vector field on 𝐷. The idea here is that each point in 𝐷 gets
assigned a vector. A picture for this is provided by a model of fluid flow
in which it point in region where fluid is flowing gets assigned a vector
giving the velocity of the flow at that particular point.
27
28 CHAPTER 3. FUNCTIONS
∙ Scalar Fields. For the case in which 𝑚 = 1 and 𝑛 > 1, every point in
𝐷 now gets assigned a scalar (a real number). An example of this in
applications would be the temperature distribution over a region in space.
Scalar fields in this course will usually be denoted by lower case letters (𝑓 ,
𝑔, etc.). The value of a scalar field
𝑓: 𝐷 →ℝ
at a point 𝑃 (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ) in 𝐷 will be denoted by
𝑓 (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ).
If 𝐷 is a region in the 𝑥𝑦–plane, we simply write
𝑓 (𝑥, 𝑦) for (𝑥, 𝑦) ∈ 𝐷.
Example 3.2.3. For any 𝑅 > 0, the open ball 𝐵𝑅 (𝑂) = {𝑦 ∈ ℝ𝑛 ∣ ∥𝑦∥ < 𝑅} is
an open set.
Proof. Let 𝑥 be an arbitrary point in 𝐵𝑅 (𝑂); then ∥𝑥∥ < 𝑅. Put 𝑟 = 𝑅−∥𝑥∥ > 0
and consider the open ball 𝐵𝑟 (𝑥). If 𝑦 ∈ 𝐵𝑟 (𝑥), then, by the triangle inequality,
𝐵𝑟 (𝑥) ⊆ 𝐵𝑅 (𝑂).
This is the expression that we will use to generalize the notion of continuity at a
point to vector valued functions on subsets of Euclidean space. We will simply
replace the absolute values by norms.
= (𝑥2 − 𝑥2𝑜 )2 + (𝑦 − 𝑦𝑜 )2 ,
which may be written as
( ) ( ) 2
𝑥 𝑥𝑜
𝐹 −𝐹 = (𝑥 + 𝑥𝑜 )2 (𝑥 − 𝑥𝑜 )2 + (𝑦 − 𝑦𝑜 )2 , (3.1)
𝑦 𝑦𝑜
after factoring.
( )
𝑥
Next, restrict to values of ∈ ℝ2 such that
𝑦
( ) ( )
𝑥 𝑥𝑜
− ⩽ 1. (3.2)
𝑦 𝑦𝑜
where we have used the triangle inequality. It follows from the last
inequality in (3.3) that
( )
𝑥𝑜
Hence, 𝐹 is continuous at . □
𝑦𝑜
First, write
𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 ) = 𝑥𝑦 − 𝑥𝑜 𝑦𝑜 = 𝑥𝑦 − 𝑥𝑜 𝑦 + 𝑥𝑜 𝑦 − 𝑥𝑜 𝑦𝑜 ,
or
𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 ) = 𝑦(𝑥 − 𝑥𝑜 ) + 𝑥𝑜 (𝑦 − 𝑦𝑜 ). (3.9)
32 CHAPTER 3. FUNCTIONS
Taking absolute values on both sides of (3.9) and applying the tri-
angle inequality yields that
∣𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 )∣ ⩽ ∣𝑦∣∣𝑥 − 𝑥𝑜 ∣ + ∣𝑥𝑜 ∣∣𝑦 − 𝑦𝑜 ∣. (3.10)
Restricting to values of (𝑥, 𝑦) such that
∥(𝑥, 𝑦) − (𝑥𝑜 , 𝑦𝑜 )∥ ⩽ 1, (3.11)
we see that
√ √
∣𝑦 − 𝑦𝑜 ∣ = (𝑦 − 𝑦𝑜 )2 ⩽ (𝑥 − 𝑥𝑜 )2 + (𝑦 − 𝑦𝑜 )2 ⩽ 1,
so that
∣𝑦∣ = ∣𝑦 − 𝑦𝑜 + 𝑦𝑜 ∣ ⩽ ∣𝑦 − 𝑦𝑜 ∣ + ∣𝑦𝑜 ∣ ⩽ 1 + ∣𝑦𝑜 ∣, (3.12)
provided that (3.11) holds. Thus, using the estimate in (3.12) in
(3.10), we obtain that, if (𝑥, 𝑦) satisfies (3.11),
∣𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 )∣ ⩽ (1 + ∣𝑦𝑜 ∣)∣𝑥 − 𝑥𝑜 ∣ + ∣𝑥𝑜 ∣∣𝑦 − 𝑦𝑜 ∣. (3.13)
Next, apply the Cauchy–Schwarz inequality to the right–hand side
of (3.13) to obtain
√ √
∣𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 )∣ ⩽ (1 + ∣𝑦𝑜 ∣)2 + 𝑥2𝑜 (𝑥 − 𝑥𝑜 )2 + (𝑦 − 𝑦𝑜 )2 ,
or
∣𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 )∣ ⩽ 𝐶𝑜 ∥(𝑥, 𝑦) − (𝑥𝑜 , 𝑦𝑜 )∥,
√
for values of (𝑥, 𝑦) within 1 of (𝑥𝑜 , 𝑦𝑜 ), where 𝐶𝑜 = (1 + ∣𝑦𝑜 ∣)2 + 𝑥2𝑜 .
We then have that, if ∥(𝑥, 𝑦) − (𝑥𝑜 , 𝑦𝑜 )∥ ⩽ 1, then
0 ⩽ ∣𝑓 (𝑥, 𝑦) − 𝑓 (𝑥𝑜 , 𝑦𝑜 )∣ ⩽ 𝐶𝑜 ∥(𝑥, 𝑦) − (𝑥𝑜 , 𝑦𝑜 )∥. (3.14)
The claim in (3.8) now follows by applying the Squeeze Theorem to
the expressions in (3.14) since the rightmost term in (3.14) goes to
0 as ∥(𝑥, 𝑦) − (𝑥𝑜 , 𝑦𝑜 )∥ → 0. □
Proposition 3.3.5. Let 𝑈 denote an open subset of ℝ𝑛 and 𝐹 : 𝑈 → ℝ𝑚 be a
vector valued function defined on 𝑈 and given by
⎛ ⎞
𝑓1 (𝑣)
⎜ 𝑓2 (𝑣) ⎟
𝐹 (𝑣) = ⎜ . ⎟ , for all 𝑣 ∈ 𝑈,
⎜ ⎟
⎝ .. ⎠
𝑓𝑚 (𝑣)
where
𝑓𝑗 : 𝑈 → ℝ, for 𝑗 = 1, 2, . . . 𝑚,
are real valued functions defined on 𝑈 . The vector valued function, 𝐹 , is con-
tinuous at 𝑢 ∈ 𝑈 if and only if each one of its components, 𝑓𝑗 , for 𝑗 = 1, 2, . . . 𝑚,
is continuous at 𝑢.
3.3. CONTINUOUS FUNCTIONS 33
if and only if ⎛ ⎞
∑𝑚
lim ⎝ ∣𝑓𝑗 (𝑣) − 𝑓𝑗 (𝑢)∣2 ⎠ = 0,
∥𝑣−𝑢∥→0
𝑗=1
if and only if
𝑚
∑
lim ∣𝑓𝑗 (𝑣) − 𝑓𝑗 (𝑢)∣2 = 0,
∥𝑣−𝑢∥→0
𝑗=1
if and only if
if and only if
for all 𝑡 in some interval (𝑎, 𝑏) of real numbers. Since the sine and cosine
functions are continuous everywhere on ℝ, it follows that the path is continuous.
Proof: Write
𝑤1𝑇 𝑣
⎛ ⎞
⎜ 𝑤2𝑇 𝑣 ⎟
𝐹 (𝑣) = ⎜ . ⎟ , for all 𝑣 ∈ ℝ𝑛 ,
⎜ ⎟
⎝ .. ⎠
𝑇
𝑣𝑚 𝑣
where 𝑤1𝑇 , 𝑤2𝑇 , . . . , 𝑤𝑚
𝑇
are the rows of the matrix representation of the function
𝐹 relative to the standard basis in ℝ𝑛 . It then follows that
⎛ ⎞
𝑓1 (𝑣)
⎜ 𝑓2 (𝑣)⎟
𝐹 (𝑣) = ⎜ . ⎟ , for all 𝑣 ∈ ℝ𝑛 ,
⎜ ⎟
⎝ .. ⎠
𝑓𝑚 (𝑣)
where
𝑓𝑗 (𝑣) = 𝑤𝑗 ⋅ 𝑣, for all 𝑣 ∈ ℝ𝑛 ,
and 𝑗 = 1, 2, . . . , 𝑚. As shown in Example 3.3.2, each 𝑓𝑗 is continuous at every
𝑢 ∈ ℝ𝑛 . It then follows from Proposition 3.3.5 that 𝐹 is continuous at every
𝑢 ∈ ℝ𝑛 .
Example 3.3.8. Define 𝑓 : ℝ𝑛 → ℝ by 𝑓 (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ) = 𝑥𝑖 , for a fixed 𝑖 in
{1, 2, . . . , 𝑛}. Show that 𝑓 is continuous on ℝ.
Solution: Observe that 𝑓 is linear. In fact, note that
𝑓 (𝑣) = 𝑒𝑖 ⋅ 𝑣, for all 𝑣 ∈ ℝ𝑛 ,
where 𝑒𝑖 is the 𝑖th vector in the standard basis of ℝ𝑛 . It follows
from the result of Example 3.3.7 that 𝑓 is continuous on ℝ𝑛 . □
Example 3.3.9 (Orthogonal Projections are Continuous). Let 𝑢
ˆ denote a unit
vector in ℝ𝑛 and define 𝑃𝑢ˆ : ℝ𝑛 → ℝ𝑛 by
𝑃𝑢ˆ (𝑣) = (𝑣 ⋅ 𝑢
ˆ)ˆ
𝑢, for all 𝑣 ∈ ℝ𝑛 .
Prove that 𝑃𝑢ˆ is continuous on ℝ𝑛 .
Solution: Observe that 𝑃𝑢ˆ is linear. In fact, for any 𝑐1 , 𝑐2 ∈ ℝ and
𝑣1 , 𝑣2 ∈ ℝ𝑛 ,
𝑃𝑢ˆ (𝑐1 𝑣1 + 𝑐2 𝑣2 ) = [(𝑐1 𝑣1 + 𝑐2 𝑣2 ) ⋅ 𝑢
ˆ]ˆ
𝑢
= (𝑐1 𝑣1 ⋅ 𝑢
ˆ + 𝑐2 𝑣2 ⋅ 𝑢
ˆ)ˆ
𝑢
= (𝑐1 𝑣1 ⋅ 𝑢 𝑢 + (𝑐2 𝑣2 ⋅ 𝑢
ˆ)ˆ ˆ)ˆ
𝑢
= 𝑐1 (𝑣1 ⋅ 𝑢 𝑢 + 𝑐2 (𝑣2 ⋅ 𝑢
ˆ)ˆ ˆ)ˆ
𝑢
Example 3.3.12. Let 𝜎 be as in the previous example, and 𝐴 = (0, 𝜋/2). Then,
Example 3.3.13. Let 𝐷 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑥2 + 𝑦 2 < 1}, the open unit disc in
ℝ2 , and 𝑓 : 𝐷′ → ℝ be given by
√
𝑓 (𝑥, 𝑦) = 1 − 𝑥2 − 𝑦 2 , for (𝑥, 𝑦) ∈ 𝐷
Solution:
𝑓 −1 (0) = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑓 (𝑥, 𝑦) = 0}.
Now, 𝑓 (𝑥, 𝑦) = 0 if and only if
√
1 − 𝑥2 − 𝑦 2 = 0
if and only if
𝑥2 + 𝑦 2 = 1.
Thus,
𝑓 −1 (0) = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑥2 + 𝑦 2 = 1},
or the unit circle around the origin in ℝ2 . □
In other words, 𝐹 (𝑦) can be made arbitrarily close to 𝐹 (𝑥) by making 𝑦 suffi-
ciently close to 𝑥.
Let 𝑉 denote an arbitrary open subset of ℝ𝑚 and consider
𝐹 −1 (𝑉 ) = {𝑥 ∈ 𝑈 ∣ 𝐹 (𝑥) ∈ 𝑉 }.
𝐵𝜀 (𝐹 (𝑥)) ⊆ 𝑉.
In other words,
𝐵𝛿 (𝑥) ⊆ 𝐹 −1 (𝑉 ).
Therefore, 𝐹 −1 (𝑉 ) is open, an so the claim is proved.
Conversely, assume that for any open subset, 𝑉 , of ℝ𝑚 , 𝐹 −1 (𝑉 ) is open. We
show that this implies that 𝐹 is continuous at any 𝑥 ∈ 𝑈 . To see this, suppose
that 𝑥 ∈ 𝑈 and let 𝜀 > 0 be arbitrary. Now, since 𝐵𝜀 (𝐹 (𝑥)), the open ball of
radius 𝜀 around 𝐹 (𝑥), is an open subset of ℝ𝑚 , it follows that
𝐹 −1 (𝐵𝜀 (𝐹 (𝑥)))
is open, by the assumption we are making in this part of the proof. Hence, since
𝑥 ∈ 𝐹 −1 (𝐵𝜀 (𝐹 (𝑥))), there exists 𝛿 > 0 such that
or
∥𝑦 − 𝑥∥ < 𝛿 implies that 𝐹 (𝑦) ∈ 𝐵𝜀 (𝐹 (𝑥)),
3.3. CONTINUOUS FUNCTIONS 37
or
∥𝑦 − 𝑥∥ < 𝛿 implies that ∥𝐹 (𝑦) − 𝐹 (𝑥)∥ < 𝜀.
Thus, given an arbitrary 𝜀 > 0, there exists 𝛿 > 0 such that
𝐹 (𝑈 ) ⊆ 𝑄.
𝑥 ∈ (𝐺 ∘ 𝐹 )−1 (𝑉 ) iff (𝐺 ∘ 𝐹 )(𝑥) ∈ 𝑉
iff 𝐺(𝐹 (𝑥)) ∈ 𝑉
iff 𝐹 (𝑥) ∈ 𝐺−1 (𝑉 )
iff 𝑥 ∈ 𝐹 −1 (𝐺−1 (𝑉 )),
so that
(𝐺 ∘ 𝐹 )−1 (𝑉 ) = 𝐹 −1 (𝐺−1 (𝑉 )).
38 CHAPTER 3. FUNCTIONS
Example 3.3.16 (Evaluating scalar fields on paths). Let (𝑎, 𝑏) denote an open
interval of real numbers and 𝜎 : (𝑎, 𝑏) → ℝ𝑛 be a path. Given a scalar field
𝑓 : ℝ𝑛 → ℝ, we can define the composition
𝑓 ∘ 𝜎 : (𝑎, 𝑏) → ℝ
by 𝑓 ∘ 𝜎(𝑡) = 𝑓 (𝜎(𝑡)) for all 𝑡 ∈ (𝑎, 𝑏). Thus, 𝑓 ∘ 𝜎 is a real valued function
of a single variable like those studied in Calculus I and II. An example of a
composition 𝑓 ∘ 𝜎 is provided by evaluating the electrostatic potential, 𝑓 , along
the path of a particle moving according to 𝜎(𝑡), where 𝑡 denotes time.
According to Proposition 3.3.15, if both 𝑓 and 𝜎 are continuous, then so is
the function 𝑓 ∘ 𝜎. Therefore, if lim 𝜎(𝑡) = 𝑥𝑜 for some 𝑡𝑜 ∈ (𝑎, 𝑏) and 𝑥𝑜 ∈ ℝ𝑛 ,
𝑡→𝑡𝑜
then
lim 𝑓 (𝜎(𝑡)) = 𝑓 (𝑥𝑜 ).
𝑡→𝑡𝑜
The point here is that, if 𝑓 is continuous at 𝑥𝑜 , the limit of 𝑓 along any con-
tinuous path that approaches 𝑥𝑜 must yield the same value of 𝑓 (𝑥𝑜 ).
In other words, taking the limit along any continuous path approaching 𝑥𝑜 as
𝑡 → 𝑡𝑜 must yield one, and only one, value.
∣𝑥∣
𝑓 (𝑥, 𝑦) = √ , for (𝑥, 𝑦) ∕= (0, 0).
𝑥2 + 𝑦 2
∣𝑥∣
lim √
(𝑥,𝑦)→(0,0) 𝑥2 + 𝑦 2
cannot exist. □
40 CHAPTER 3. FUNCTIONS
Chapter 4
Differentiability
𝑓 (𝑥) − 𝑓 (𝑎)
lim
𝑥→𝑎 𝑥−𝑎
exists. If this limit exists, we denote it by 𝑓 ′ (𝑎) and call it the derivative of 𝑓
at 𝑎. We then have that
𝑓 (𝑥) − 𝑓 (𝑎)
lim = 𝑓 ′ (𝑎).
𝑥→𝑎 𝑥−𝑎
The last expression is equivalent to
𝑓 (𝑥) − 𝑓 (𝑎)
lim − 𝑓 ′ (𝑎) = 0,
𝑥→𝑎 𝑥−𝑎
∣𝐸𝑎 (𝑥 − 𝑎)∣
lim = 0;
𝑥→𝑎 ∣𝑥 − 𝑎∣
41
42 CHAPTER 4. DIFFERENTIABILITY
by (4.6). Similarly,
∥𝐸(𝑡ˆ
𝑢)∥
lim = 0.
∣𝑡∣→0 ∣𝑡∣
Thus, letting 𝑡 → 0 in (4.8) we get that
𝑇 (ˆ
𝑢) = 𝑇 (ˆ
𝑢).
Hence 𝑇 agrees with 𝑇𝑥 on any unit vector 𝑢 ˆ. Therefore, 𝑇 and 𝑇𝑥 agree on the
standard basis {𝑒1 , 𝑒2 , . . . , 𝑒𝑛 } of ℝ𝑛 . Consequently, since 𝑇 and 𝑇𝑥 are linear
where
∥𝐸𝑥 (𝑤)∥
lim = 0.
∥𝑤∥→0 ∥𝑤∥
∣𝐸𝑥 (𝑤)∣
lim = 0. (4.10)
∥𝑤∥→0 ∥𝑤∥
Now, since 𝐷𝑓 (𝑥) is a linear map from ℝ𝑛 to ℝ, there exists an 𝑛–row vector
𝑣 = [ 𝑎1 𝑎2 ⋅⋅⋅ 𝑎𝑛 ]
such that
𝐷𝑓 (𝑥)𝑤 = 𝑣 ⋅ 𝑤 for all 𝑤 ∈ ℝ𝑛 ; (4.11)
that is, 𝐷𝑓 (𝑥)𝑤 is the dot–product of 𝑣 an 𝑤. We would like to know what the
differentiability of 𝑓 implies about the components of the vector 𝑣.
4.3. EXAMPLE: DIFFERENTIABLE SCALAR FIELDS 45
𝑓 (𝑥 + 𝑡ˆ
𝑒𝑗 ) = 𝑓 (𝑥) + 𝐷𝑓 (𝑥)(𝑡ˆ
𝑒𝑗 ) + 𝐸𝑥 (𝑡ˆ
𝑒𝑗 ). (4.12)
Using the linearity of 𝐷𝑓 (𝑥) and (4.11) we get from (4.12) that
𝑒𝑗 ) − 𝑓 (𝑥)
𝑓 (𝑥 + 𝑡ˆ 𝐸𝑥 (𝑡ˆ𝑒𝑗 )
= 𝑎𝑗 + . (4.13)
𝑡 𝑡
It follows from (4.10) that
∣𝐸𝑥 (𝑡ˆ
𝑒𝑗 )∣ ∣𝐸𝑥 (𝑡ˆ
𝑒𝑗 )∣
lim = lim = 0,
𝑡→0 ∣𝑡∣ ∣𝑡∣→0 ∥𝑡ˆ
𝑒𝑗 ∥
𝑒𝑗 ) − 𝑓 (𝑥)
𝑓 (𝑥 + 𝑡ˆ
lim = 𝑎𝑗 . (4.14)
𝑡→0 𝑡
Definition 4.3.1 (Partial Derivatives). Let 𝑈 be an open subset of ℝ𝑛 ,
𝑓: 𝑈 →ℝ
𝑒𝑗 ) − 𝑓 (𝑥)
𝑓 (𝑥 + 𝑡ˆ
lim
𝑡→0 𝑡
exists, we call it the partial derivative of 𝑓 at 𝑥 with respect to 𝑥𝑗 and denote
∂𝑓
it by (𝑥).
∂𝑥𝑗
The argument leading up to equation (4.14) then shows that if the scalar
field 𝑓 : 𝑈 → ℝ is differentiable at 𝑥 ∈ 𝑈 , then its partial derivatives at 𝑥 exist
and they are the components of the matrix representation of the linear map
𝐷𝑓 (𝑥) : ℝ𝑛 → ℝ with respect to the standard basis in ℝ𝑛 :
[ ]
∂𝑓 ∂𝑓 ∂𝑓
[𝐷𝑓 (𝑥)] = (𝑥) (𝑥) ⋅ ⋅ ⋅ (𝑥) .
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
∂𝑓 𝑓 (𝑥 + 𝑡, 𝑦) − 𝑓 (𝑥, 𝑦)
(𝑥, 𝑦) = lim .
∂𝑥 𝑡→0 𝑡
Thus, we compute the rate of change of 𝑓 as 𝑥 changes while 𝑦 is
fixed. For the case in which (𝑥, 𝑦) ∕= (0, 0), we may compute ∂𝑓 /∂𝑥
as follows:
1
⎛ ⎞
∂𝑓 − 2 2⎟
(𝑥, 𝑦) = ∂𝑥 ⎝𝑒 𝑥 + 𝑦 ⎠
∂ ⎜
∂𝑥
1
− 2 2 ∂
(
1
)
= 𝑒 𝑥 +𝑦 ⋅ − 2
∂𝑥 𝑥 + 𝑦2
1
− 2 2 2𝑥
= 𝑒 𝑥 +𝑦 ⋅ 2
(𝑥 + 𝑦 2 )2
1
2𝑥 − 2
= ⋅𝑒 𝑥 + 𝑦2 .
(𝑥2 + 𝑦 2 )2
∂𝑓 𝑓 (𝑡, 0) − 𝑓 (0, 0)
(0, 0) = lim
∂𝑥 𝑡→0 𝑡
1
− 2
𝑒 𝑡
= lim
𝑡→0 𝑡
1/𝑡
= lim .
𝑡→0 1/𝑡2
𝑒
Applying L’Hospital’s Rule we then have that
∂𝑓 1/𝑡2
(0, 0) = lim 2
∂𝑥 𝑡→0
2/𝑡3 𝑒1/𝑡
1 𝑡
= lim
2 𝑡→0 𝑒1/𝑡2
= 0.
∂𝑓
Similarly, (0, 0) = 0. It then follows that
∂𝑦
where
∣𝐸(𝑥, 𝑦)∣
lim √ = 0,
(𝑥,𝑦)→(0,0) 𝑥2 + 𝑦 2
and 𝑇 is the zero linear transformation from ℝ2 to ℝ.
In this case
1
−
𝐸(𝑥, 𝑦) = 𝑒 𝑥2 + 𝑦2 if (𝑥, 𝑦) ∕= (0, 0).
Thus, for (𝑥, 𝑦) ∕= (0, 0),
1 1
− 2 2 −
∣𝐸(𝑥, 𝑦)∣ 𝑒 𝑥 +𝑦 𝑒 𝑢2
√ = √ = ,
𝑥2 + 𝑦 2 𝑥2 + 𝑦 2 𝑢
√
where we have set 𝑢 = 𝑥2 + 𝑦 2 . Thus,
1
−
∣𝐸(𝑥, 𝑦)∣ 𝑒 𝑢2
lim √ = lim = 0,
(𝑥,𝑦)→(0,0) 𝑥2 + 𝑦 2 𝑢→0 𝑢
by the same calculation involving L’Hospital’s Rule that was used to
compute ∂𝑓 /∂𝑥 at (0, 0). Consequently, 𝑓 is differentiable at (0, 0)
and its derivative is the zero map. □
We have seen that if a scalar field 𝑓 : 𝑈 → ℝ is differentiable at 𝑥 ∈ 𝑢, then
𝑓 (𝑥 + 𝑤) = 𝑓 (𝑥) + ∇𝑓 (𝑥) ⋅ 𝑤 + 𝐸𝑥 (𝑤)
for all 𝑤 ∈ ℝ𝑛 with sufficiently small norm, ∥𝑤∥, where ∇𝑓 (𝑥) is the gradient
of 𝑓 at 𝑥 ∈ 𝑈 , and
∣𝐸𝑥 (𝑤)∣
lim = 0.
∥𝑤∥→0 ∥𝑤∥
Applying this to the case where 𝑤 = 𝑡ˆ
𝑢, for a unit vector 𝑢
ˆ, we get that
𝑢) − 𝑓 (𝑥) = 𝑡∇𝑓 (𝑥) ⋅ 𝑢
𝑓 (𝑥 + 𝑡ˆ ˆ + 𝐸𝑥 (𝑡ˆ
𝑢)
for 𝑡 ∈ ℝ sufficiently close to 0. Dividing by 𝑡 ∕= 0 and letting 𝑡 → 0 leads to
𝑢) − 𝑓 (𝑥)
𝑓 (𝑥 + 𝑡ˆ
lim = ∇𝑓 (𝑥) ⋅ 𝑢
ˆ,
𝑡→0 𝑡
where we have used (4.10).
Definition 4.3.4 (Directional Derivatives). Let 𝑓 : 𝑈 → ℝ denote a scalar field
defined on an open subset 𝑈 of ℝ𝑛 , and let 𝑢
ˆ be a unit vector in ℝ𝑛 . If the limit
𝑢) − 𝑓 (𝑥)
𝑓 (𝑥 + 𝑡ˆ
lim
𝑡→0 𝑡
exists, we call it the directional derivative of 𝑓 at 𝑥 in the direction of the unit
vector 𝑢ˆ. We denote it by 𝐷𝑢ˆ 𝑓 (𝑥).
4.4. EXAMPLE: DIFFERENTIABLE PATHS 49
We have then shown that if the scalar field 𝑓 is differentiable at 𝑥, then its
directional derivative at 𝑥 in the direction of a unit vector 𝑢
ˆ is given by
that is, the dot–product of the gradient of 𝑓 at 𝑥 with the unit vector 𝑢 ˆ. In
other words, the directional derivative on 𝑓 at 𝑥 in the direction of a unit vector
ˆ is the component of the orthogonal projection of ∇𝑓 (𝑥) along the direction of
𝑢
𝑢
ˆ.
where
∥𝐸𝑡 (ℎ)∥
lim = 0. (4.16)
ℎ→0 ∣ℎ∣
v(𝑡) = 𝐷𝜎(𝑡)(1);
that is, v(𝑡) is the image of the real number 1 under the linear transforma-
tion 𝐷𝜎(𝑡).
(b) Write 𝜎(𝑡) = (𝑥1 (𝑡), 𝑥2 (𝑡), . . . , 𝑥𝑛 (𝑡)) for all 𝑡 ∈ 𝐼. Show that if 𝜎 : 𝐼 → ℝ𝑛
is differentiable at 𝑡 ∈ 𝐼 and v = 𝐷𝜎(𝑡)(1), then each function 𝑥𝑗 : 𝐼 → ℝ,
for 𝑗 = 1, 2, . . . , 𝑛, is differentiable at 𝑡, and
is called the the tangent line to the path 𝜎(𝑡) and the point 𝜎(𝑡𝑜 ).
Example 4.4.3. Give the tangent line to the path
when 𝑡𝑜 = 𝜋/4.
where 𝜎 ′ (𝑡) = (− sin 𝑡, 1, cos 𝑡); so that, for 𝑡𝑜 = 𝜋/4, we get that
(√ √ ) ( ( √ √ )
2 𝜋 2 𝜋) 2 2
𝑟(𝑡) = , , + 𝑡− − , 1,
2 4 2 4 2 2
for 𝑡 ∈ ℝ.
Writing (𝑥, 𝑦, 𝑧) for the vector 𝑟(𝑡), we obtain the parametric equa-
tions for the tangent line:
⎧ √ √ (
2 2 𝜋
)
⎨𝑥 = 2 − 2 𝑡 − 4
𝜋
𝑦=√ 4 +𝑡 √ (
𝑧 = 22 + 22 𝑡 − 𝜋4
⎩ )
where
∣𝐸𝑗 (𝑡, ℎ)∣
lim = 0, for all 𝑗 = 1, 2, . . . 𝑛. (4.20)
ℎ→0 ∣ℎ∣
It follows from (4.19) that
Putting
𝑥′1 (𝑡)
⎛ ⎞
⎜ 𝑥′2 (𝑡) ⎟
𝜎(𝑡) = ⎜ . ⎟ , (4.22)
⎜ ⎟
⎝ .. ⎠
𝑥′𝑛 (𝑡)
we obtain from the equations in (4.21) that
where 𝐸1 (𝑡, ℎ), 𝐸2 (𝑡, ℎ), . . . , 𝐸𝑛 (𝑡, ℎ) are given in (4.19) and satisfy (4.20). It
then follows that, for ℎ ∕= 0 and ∣ℎ∣ small enough,
⎛ ⎞
𝐸1 (𝑡, ℎ)/ℎ
1 ⎜ 𝐸2 (𝑡, ℎ)/ℎ ⎟
(𝜎(𝑡 + ℎ) − 𝜎(𝑡) − ℎ𝜎 ′ (𝑡)) = ⎜ ⎟.
⎜ ⎟
..
ℎ ⎝ . ⎠
𝐸𝑛 (𝑡, ℎ)/ℎ
We prove that
∣𝐸(ℎ, 𝑘)∣
lim √ =0 (4.24)
(ℎ,𝑘)→(0,0) ℎ2 + 𝑘 2
54 CHAPTER 4. DIFFERENTIABILITY
Assume that ℎ > 0 and 𝑘 > 0 (the other cases can be treated in an analogous
manner). By the mean value theorem, there are real numbers 𝜃 and 𝜂 such that
0 < 𝜃 < 1 and 0 < 𝜂 < 1 and
∂𝑓
𝑓 (𝑥 + ℎ, 𝑦 + 𝑘) − 𝑓 (𝑥, 𝑦 + 𝑘) = (𝑥 + 𝜃ℎ, 𝑦 + 𝑘) ⋅ ℎ,
∂𝑥
and
∂𝑓
𝑓 (𝑥, 𝑦 + 𝑘) − 𝑓 (𝑥, 𝑦) = (𝑥, 𝑦 + 𝜂𝑘) ⋅ 𝑘.
∂𝑦
Consequently,
∂𝑓 ∂𝑓
𝑓 (𝑥 + ℎ, 𝑦 + 𝑘) − 𝑓 (𝑥, 𝑦) = (𝑥 + 𝜃ℎ, 𝑦 + 𝑘) ⋅ ℎ + (𝑥, 𝑦 + 𝜂𝑘) ⋅ 𝑘.
∂𝑥 ∂𝑦
Thus, 𝐸(ℎ, 𝑘) is the dot product of the vector 𝑣(ℎ, 𝑘), given by
( )
∂𝑓 ∂𝑓 ∂𝑓 ∂𝑓
𝑣(ℎ, 𝑘) = (𝑥 + 𝜃ℎ, 𝑦 + 𝑘) − (𝑥, 𝑦), (𝑥, 𝑦 + 𝜂𝑘) − (𝑥, 𝑦) ,
∂𝑥 ∂𝑥 ∂𝑦 ∂𝑥
∣𝐸(ℎ, 𝑘)∣
√ ⩽ ∥𝑣(ℎ, 𝑘)∥, (4.25)
ℎ2 + 𝑘 2
where
√( )2 ( )2
∂𝑓 ∂𝑓 ∂𝑓 ∂𝑓
∥𝑣(ℎ, 𝑘)∥ = (𝑥 + 𝜃ℎ, 𝑦 + 𝑘) − (𝑥, 𝑦) + (𝑥, 𝑦 + 𝜂𝑘) − (𝑥, 𝑦)
∂𝑥 ∂𝑥 ∂𝑦 ∂𝑥
∣𝐸(ℎ, 𝑘)∣
lim √ = 0,
(ℎ,𝑘)→(0,0) ℎ2 + 𝑘 2
which is (4.24). This shows that 𝑓 is differentiable at (𝑥, 𝑦). Since (𝑥, 𝑦) was
arbitrary, the result follows.
4.5. SUFFICIENT CONDITION FOR DIFFERENTIABILITY 55
∂𝑓𝑖
(𝑥) 𝑖 = 1, 2, . . . , 𝑚; 𝑗 = 1, 2, . . . , 𝑛,
∂𝑥𝑗
are continuous on 𝑈 .
is differentiable at (0, 0); however, the partial derivatives are not continuous at
the origin (This is shown in Problem 9 of Assignment #5).
𝐷𝐹 (𝑥) : ℝ𝑛 → ℝ𝑚
56 CHAPTER 4. DIFFERENTIABILITY
is given by
∂𝑓1 ∂𝑓1 ∂𝑓1
⎛ ⎞
⎜ ∂𝑥1 (𝑥) (𝑥) ⋅⋅⋅ (𝑥) ⎟
⎜ ∂𝑥2 ∂𝑥𝑛 ⎟
⎜ ⎟
⎜ ∂𝑓 ∂𝑓2 ∂𝑓2 ⎟
⎜ 2
(𝑥) (𝑥) ⋅ ⋅ ⋅ (𝑥) ⎟ .
⎟
⎜ ∂𝑥1
⎜
∂𝑥2 ∂𝑥𝑛 ⎟ (4.26)
⎜ .
.. .. .. .. ⎟
⎜
⎜ . . . ⎟
⎟
⎝ ∂𝑓𝑚 ∂𝑓𝑚 ∂𝑓𝑚 ⎠
(𝑥) (𝑥) ⋅ ⋅ ⋅ (𝑥)
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
The matrix of partial derivative of the components of 𝐹 in equation (4.26) is
called the Jacobian matrix of the map 𝐹 at 𝑥. It is the matrix that represents
the derivative map 𝐷𝐹 (𝑥) : ℝ𝑛 → ℝ𝑚 with respect to the standard bases in
ℝ𝑛 and ℝ𝑚 . We will therefore denote it by 𝐷𝐹 (𝑥). Hence, 𝐷𝐹 (𝑥)𝑤 can be
understood as matrix multiplication of the Jacobian matrix of 𝐹 at 𝑥 by the
column vector 𝑤. If 𝑚 = 𝑛, then the determinant of the square matrix 𝐷𝐹 (𝑥)
is called the Jacobian determinant of 𝐹 at 𝑥, and is denoted by the symbols
∂(𝑓1 , 𝑓2 , . . . , 𝑓𝑛 )
𝐽𝐹 (𝑥) or . We then have that
∂(𝑥1 , 𝑥2 , . . . , 𝑥𝑛 )
∂(𝑓1 , 𝑓2 , . . . , 𝑓𝑛 )
𝐽𝐹 (𝑥) = = det 𝐷𝐹 (𝑥).
∂(𝑥1 , 𝑥2 , . . . , 𝑥𝑛 )
𝐽𝐹 (𝑥, 𝑦) = 4(𝑥2 + 𝑦 2 ).
𝐺 ∘ 𝐹 : 𝑈 → ℝ𝑘
where
∥𝐸𝐹 (𝑤)∥
lim = 0. (4.28)
∥𝑤∥→0 ∥𝑤∥
Similarly, for 𝑣 ∈ ℝ𝑚 with ∥𝑣∥ sufficiently small,
where
∥𝐸𝐺 (𝑣)∥
lim = 0. (4.30)
∥𝑣∥→0 ∥𝑣∥
It then follows from (4.27) that, for 𝑤 ∈ ℝ𝑛 with ∥𝑤∥ sufficiently small,
where v
u𝑚 ∑ 𝑛 (
u∑ )2
∂𝑓𝑖
∥𝐷𝐹 (𝑥)∥ = ⎷ (𝑥) ;
𝑖=1 𝑗=1
∂𝑥𝑗
so that, by virtue of (4.28), we can make ∥𝑣∥ small by making ∥𝑤∥ small. It
then follows from (4.29) and (4.31) that
Put
𝐸(𝑤) = 𝐷𝐺(𝑦)𝐸𝐹 (𝑤) + 𝐸𝐺 (𝑣) (4.35)
𝑛
for 𝑤 ∈ ℝ and 𝑣 as given in (4.32). The differentiability of 𝐺 ∘ 𝐹 at 𝑥 will then
follow from (4.34) if we can prove that
∥𝐸(𝑤)∥
lim = 0. (4.36)
∥𝑤∥→0 ∥𝑤∥
This will also prove that
To prove (4.36), take the norm of 𝐸(𝑤) defined in (4.35), apply the triangle and
Cauchy–Schwarz inequalities, and divide by ∥𝑤∥ to get that
The proof of (4.36) will then follow from this last estimate, (4.28), (4.30), (4.37)
and the Squeeze Theorem. This completes the proof of the Chain Rule.
where ( )
𝑥(𝑢, 𝑣)
Φ(𝑢, 𝑣) = , for (𝑢, 𝑣) ∈ 𝑄,
𝑦(𝑢, 𝑣)
we have that
𝐷𝑔(𝑢, 𝑣) = 𝐷𝑓 (𝑥(𝑢, 𝑣), 𝑦(𝑢, 𝑣))𝐷Φ(𝑢, 𝑣).
Writing this in terms of Jacobian matrices we get
⎛ ⎞
∂𝑥 ∂𝑥
∂𝑢 ∂𝑣 ⎟
( ) ( )⎜
∂𝑔 ∂𝑔 ∂𝑓 ∂𝑓 ⎜
= ⎟,
⎟
∂𝑥 ∂𝑦 ⎝ ∂𝑦
⎜
∂𝑢 ∂𝑣 ∂𝑦 ⎠
∂𝑢 ∂𝑣
4.6. DERIVATIVES OF COMPOSITIONS 59
∂𝑔 ∂𝑓 ∂𝑥 ∂𝑓 ∂𝑦
= +
∂𝑢 ∂𝑥 ∂𝑢 ∂𝑦 ∂𝑢
and
∂𝑔 ∂𝑓 ∂𝑥 ∂𝑓 ∂𝑦
= + .
∂𝑣 ∂𝑥 ∂𝑣 ∂𝑦 ∂𝑣
In the previous example, if Φ : 𝑄 → ℝ2 is a one–to–one map, then Φ is called
a change of variable map. Writing Φ in terms of a its components we have
𝑥 = 𝑥(𝑢, 𝑣)
𝑦 = 𝑦(𝑢, 𝑣),
𝑥 = 𝑟 cos 𝜃
𝑦 = 𝑟 sin 𝜃,
and
∂𝑓 ∂𝑓 ∂𝑥 ∂𝑓 ∂𝑦
= +
∂𝜃 ∂𝑥 ∂𝜃 ∂𝑦 ∂𝜃
give the partial derivatives of 𝑓 with respect to the polar variables 𝑟 and 𝜃 in
terms of the partial derivatives of 𝑓 with respect to the Cartesian coordinates
𝑥 and 𝑦 and the derivative of the change of variables map
( )
𝑟 cos 𝜃
Φ(𝑟, 𝜃) = .
𝑟 sin 𝜃
d
𝑓 (𝜎(𝑡)) = ∇𝑓 (𝜎(𝑡)) ⋅ 𝜎 ′ (𝑡) for all 𝑡 ∈ 𝐼.
d𝑡
Integration
for a real valued function, 𝑓 , defined on a closed and bounded interval [𝑎, 𝑏].
We begin by defining integrals of scalar fields over curves in ℝ𝑛 which can be
parametrized by 𝐶 1 paths.
61
62 CHAPTER 5. INTEGRATION
𝑦
r
(cos 𝑡, sin 𝑡)
1 𝑥
𝜎(𝑡1 ) = 𝜎(𝑡2 )
and so
cos(𝑡1 ) = cos(𝑡2 ).
Since cos is one–to–one on [0, 𝜋/2], it follows that
𝑡1 = 𝑡2 ,
𝐶 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑥2 + 𝑦 2 = 1},
64 CHAPTER 5. INTEGRATION
𝜎(𝑡𝑁 )
r
r
H
𝜎(𝑡H
2) H 𝜎(𝑡𝑖 )
HH ( ((r
HHr(((
𝜎(𝑡𝑖−1 )
r
𝜎(𝑡1 )
r
𝜎(𝑡𝑜 )
∥𝐸𝑖 (ℎ)∥
lim = 0,
ℎ→0 ∣ℎ∣
is a Riemann sum for the function ∥𝜎 ′ (𝑡)∥ over the interval [𝑎, 𝑏]. Now, since
we are assuming the 𝜎 is of class 𝐶 1 , it follows that the map 𝑡 7→ ∥𝜎 ′ (𝑡)∥ is
66 CHAPTER 5. INTEGRATION
continuous on [𝑎, 𝑏]. Thus, a theorem from analysis guarantees that the sums
𝑁
∑
∥𝜎 ′ (𝑡𝑖−1 )∥(𝑡𝑖 − 𝑡𝑖−1 )
𝑖=1
converge as 𝑁 → ∞ while
The limit will be the Riemann integral of ∥𝜎 ′ (𝑡)∥ over the interval [𝑎, 𝑏]. Thus,
it makes sense to define ∫ 𝑏
ℓ(𝐶) = ∥𝜎 ′ (𝑡)∥d𝑡.
𝑎
We next see that we will always get the same value of the integral for any
𝐶 1 parametrization of 𝜎.
Let 𝛾(𝑡) = 𝜎(ℎ(𝑡)), for all 𝑡 ∈ [𝑐, 𝑑], be reparametrization of 𝜎 : [𝑎, 𝑏] → ℝ𝑛 ;
that is, ℎ is a one–to–one, differentiable function from [𝑐, 𝑑] to [𝑎, 𝑏] with ℎ′ (𝑡) > 0
for all 𝑡 ∈ (𝑐, 𝑑). We consider the integral
∫ 𝑑
∥𝛾 ′ (𝑡)∥d𝑡.
𝑐
𝑑
𝛾 ′ (𝑡) = [𝜎(ℎ(𝑡))] = ℎ′ (𝑡)𝜎 ′ (ℎ(𝑡)).
𝑑𝑡
We then have that
∫ 𝑑 ∫ 𝑑
′
∥𝛾 (𝑡)∥d𝑡 = ∥ℎ′ (𝑡)𝜎 ′ (ℎ(𝑡))∥d𝑡
𝑐 𝑐
∫ 𝑑
= ∥𝜎 ′ (ℎ(𝑡))∥ ∣ℎ′ (𝑡)∣d𝑡
𝑐
∫ 𝑑
= ∥𝜎 ′ (ℎ(𝑡))∥ ℎ′ (𝑡)d𝑡,
𝑐
since ℎ (𝑡) > 0. Next, make the change of variables 𝜏 = ℎ(𝑡). Then, d𝜏 = ℎ′ (𝑡)d𝑡
′
and ∫ 𝑑 ∫ 𝑏
′ ′
∥𝜎 (ℎ(𝑡))∥ℎ (𝑡)d𝑡 = ∥𝜎 ′ (𝜏 )∥d𝜏.
𝑐 𝑎
as follows: ∫ ∫ 𝑏
𝑓= 𝑓 (𝜎(𝑡))∥𝜎 ′ (𝑡)∥d𝑡, (5.1)
𝐶 𝑎
Let 𝐶 denote the image of 𝜎. Let 𝑓 (𝑥, 𝑦) denote the linear mass
density of the wire. Then, 𝑓 (𝑥, 𝑦) = 𝑘∣𝑥∣ for some constant of pro-
portionality 𝑘. It then follows that the mass of the wire is
∫ ∫ 1
𝑀= 𝑓= 𝑘∣𝑡∣∥𝜎 ′ (𝑡)∥d𝑡,
𝐶 −1
where
𝜎 ′ (𝑡) = (1, 2𝑡),
68 CHAPTER 5. INTEGRATION
so that √
∥𝜎 ′ (𝑡)∥ = 1 + 4𝑡2 .
Hence, by the symmetry of the wire with respect to the 𝑦 axis
∫ ∫ 1 √
𝑀= 𝑓 =2 𝑘𝑡 1 + 4𝑡2 d𝑡.
𝐶 0
𝑘 √
𝑀= (5 5 − 1).
6
□
∫
The definition of 𝑓 given in (5.1) is based on a choice of parametrization,
𝐶 ∫
𝜎 : [𝑎, 𝑏] → ℝ𝑛 , for 𝐶. Thus, in order to see that 𝑓 is well defined, we need
∫ 𝐶
For the case in which 𝛾 is a reparametrization of 𝜎; that is, the case in which
𝛾(𝑡) = 𝜎(ℎ(𝑡)), for all 𝑡 ∈ [𝑐, 𝑑], where ℎ is a one–to–one, differentiable function
from [𝑐, 𝑑] to [𝑎, 𝑏] with ℎ′ (𝑡) > 0 for all 𝑡 ∈ (𝑐, 𝑑). We see that (5.2) follows
from the Chain Rule and the change of variables: 𝜏 = ℎ(𝑡), for 𝑡 ∈ [𝑐, 𝑑]. In fact
we have
𝑑
𝛾 ′ (𝑡) = [𝜎(ℎ(𝑡))] = ℎ′ (𝑡)𝜎 ′ (ℎ(𝑡)),
𝑑𝑡
so that
∫ 𝑑 ∫ 𝑑
𝑓 (𝛾(𝑡))∥𝛾 ′ (𝑡)∥d𝑡 = 𝑓 (𝜎(ℎ(𝑡)))∥∥𝜎 ′ (ℎ(𝑡))∥ ℎ′ (𝑡)d𝑡,
𝑐 𝑐
which is (5.2) for the case in which one of the paths is reparametrization of the
other. Finally, using the results of Appendix B in this notes, we see that (5.2)
holds for any two parametrizations, 𝜎 : [𝑎, 𝑏] → ℝ𝑛 and 𝜎 : [𝑐, 𝑑] → ℝ𝑛 , of the
𝐶 1 simple curve, 𝐶.
5.2. LINE INTEGRALS 69
𝜎 : [𝑎, 𝑏] → ℝ𝑛 .
We have seen that the vector 𝜎 ′ (𝑡) gives the tangent direction to the path at
𝜎(𝑡). The vector
1
𝑇 (𝑡) = 𝜎 ′ (𝑡)
∥𝜎 ′ (𝑡)∥
is, therefore, a unit tangent vector to the path. The tangential component of
the of the vector field, 𝐹 , is then given by the dot product of 𝐹 and 𝑇 :
𝐹 ⋅ 𝑇.
∫ 2𝜋
= (sin2 𝑡 + cos2 𝑡)d𝑡
0
= 2𝜋.
□
Let
𝐹 (𝑥, 𝑦, 𝑧) = 𝑃 (𝑥, 𝑦, 𝑧) ˆ𝑖 + 𝑄(𝑥, 𝑦, 𝑧) ˆ
𝑗
denote a vector filed defined in a region 𝑈 of ℝ2 , where 𝑃 and 𝑄 are continuous
scalar fields defined on 𝑈 . Let
∫ 𝑏
= (𝑃 (𝑥(𝑡), 𝑦(𝑡))𝑥′ (𝑡)d𝑡 + 𝑄(𝑥(𝑡), 𝑦(𝑡))𝑦 ′ (𝑡)d𝑡)
𝑎
Next, use the notation d𝑥 = 𝑥′ (𝑡)d𝑡 and d𝑦 = 𝑦 ′ (𝑡)d𝑡 for the differentials of 𝑥
and 𝑦, respectively, to re–write the line integral as
∫ ∫
𝐹 ⋅ 𝑇 d𝑠 = 𝑃 d𝑥 + 𝑄d𝑦. (5.4)
𝐶 𝐶
Equation (5.4) suggests another way to evaluate the line integral of a 2–dimensional
vector field on a plane curve.
∫
Example 5.2.2. Evaluate the line integral −𝑦d𝑥 + (𝑥 − 1)d𝑦, where 𝐶 is
𝐶
the simple closed curve made up of the line segment from (−1, 0) to (1, 0) and
the top portion of the unit circle traversed in the counterclockwise direction (see
picture in Figure 5.2.4).
𝑦
(0, 1)
JJ 𝐶
]
2
-
(−1, 0) 𝐶1 (1, 0) 𝑥
(i) 𝐶1 : the directed line segment from (−1, 0) to (1, 0), and
(ii) 𝐶2 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑥2 + 𝑦 2 = 1, 𝑦 ⩾ 0}; the top portion of the
unit circle in ℝ2 traversed in the counterclockwise sense.
Then,
∫ ∫ ∫
−𝑦d𝑥 + (𝑥 − 1)d𝑦 = −𝑦d𝑥 + (𝑥 − 1)d𝑦 + −𝑦d𝑥 + (𝑥 − 1)d𝑦.
𝐶 𝐶1 𝐶2
𝜋
= [𝑡 − sin 𝑡]0
= 𝜋.
It then follows that
∫
−𝑦d𝑥 + (𝑥 − 1)d𝑦 = 𝜋.
𝐶
□
72 CHAPTER 5. INTEGRATION
We can obtain an analogous equation to that in (5.4) for the case of a three
dimensional field
𝐹 = 𝑃 ˆ𝑖 + 𝑄 ˆ𝑗+𝑅 ˆ 𝑘,
where 𝑃 , 𝑄 and 𝑅 are scalar fields defined in some region 𝑈 of ℝ3 which contains
the simple curve 𝐶:
∫ ∫
𝐹 ⋅ 𝑇 d𝑠 = 𝑃 d𝑥 + 𝑄d𝑦 + 𝑅d𝑧. (5.5)
𝐶 𝐶
∫ 1
= ∇𝑓 (𝜎(𝑡)) ⋅ 𝜎 ′ (𝑡)d𝑡
0
∫ 1
d
= (𝑓 (𝜎(𝑡))) d𝑡
0 d𝑡
= 𝑓 (𝜎(1)) − 𝑓 (𝜎(0))
= 𝑓 (𝑥1 ) − 𝑓 (𝑥𝑜 ).
Thus, the line integral of 𝐹 = ∇𝑓 on a curve 𝐶 is determined by the values of
𝑓 at the endpoints of the curve.
A field 𝐹 with the property that 𝐹 = ∇𝑓 , for a 𝐶 1 scalar field, 𝑓 , is called
a gradient field, and 𝑓 is called a potential for the field 𝐹 .
Example 5.3.1 (Gravitational Potential). According to Newton’s Law of Uni-
versal Gravitation, the earth exerts a gravitational pull on an object of mass 𝑚
at a point (𝑥, 𝑦, 𝑧) above the surface of the earth, which is at a distance of
√
𝑟 = 𝑥2 + 𝑦 2 + 𝑧 2
from the center of the earth (located at the origin of three dimensional space),
an is given by
𝑘𝑚
𝐹 (𝑥, 𝑦, 𝑧) = − 2 𝑟ˆ, (5.6)
𝑟
where 𝑟ˆ is a unit vector in the direction of the vector ⃗𝑟 = 𝑥 ˆ𝑖 + 𝑦 ˆ
𝑗+𝑧 ˆ
𝑘. The
minus sign indicates that the force is directed towards the center of the earth.
Show that the field 𝐹 is a gradient field.
5.4. FLUX ACROSS PLANE CURVES 73
𝑘𝑚 √
𝑓 (𝑟) = and 𝑟 = 𝑥2 + 𝑦 2 + 𝑧 2 ∕= 0. (5.7)
𝑟
To see why this is so, use the Chain Rule to compute
∂𝑓 ∂𝑟 𝑘𝑚 𝑥
= 𝑓 ′ (𝑟) =− 2 .
∂𝑥 ∂𝑥 𝑟 𝑟
Similarly,
∂𝑓 𝑘𝑚 𝑦 ∂𝑓 𝑘𝑚 𝑧
= − 2 , and =− 2 .
∂𝑦 𝑟 𝑟 ∂𝑧 𝑟 𝑟
It then follows that
∂𝑓 ˆ ∂𝑓 ˆ ∂𝑓 ˆ
∇𝑓 = 𝑖+ 𝑗+ 𝑘
∂𝑥 ∂𝑦 ∂𝑧
𝑘𝑚 𝑥 ˆ 𝑘𝑚 𝑦 ˆ 𝑘𝑚 𝑧 ˆ
= − 𝑖− 2 𝑗− 2 𝑘
𝑟2 𝑟 𝑟 𝑟 𝑟 𝑟
𝑘𝑚 ( 𝑥 ˆ 𝑦 ˆ 𝑧 ˆ)
= − 2 𝑖+ 𝑗+ 𝑘
𝑟 𝑟 𝑟 𝑟
𝑘𝑚 1 ( ˆ )
= − 2 𝑥 𝑖+𝑦 ˆ𝑗+𝑧 ˆ
𝑘
𝑟 𝑟
𝑘𝑚
= − 𝑟ˆ,
𝑟2
which is the vector field 𝐹 defined in (5.6). □
It follows from the fact that the Newtonian gravitational field 𝐹 defined in
(5.6) is a gradient field that the line integral of 𝐹 along any curve in ℝ3 , which
does not go through the origin, connecting ⃗𝑟𝑜 = (𝑥𝑜 , 𝑦𝑜 , 𝑧𝑜 ) to ⃗𝑟1 = (𝑥1 , 𝑦1 , 𝑧1 ),
is given by
∫
𝑘𝑚 𝑘𝑚
𝐹 ⋅ 𝑇 d𝑠 = 𝑓 (𝑥1 , 𝑦1 , 𝑧1 ) − 𝑓 (𝑥𝑜 , 𝑦𝑜 , 𝑧𝑜 ) = − ,
𝐶 𝑟1 𝑟𝑜
√ √
where 𝑟𝑜 = 𝑥2𝑜 + 𝑦𝑜2 + 𝑧𝑜2 and 𝑟1 = 𝑥21 + 𝑦12 + 𝑧12 . The function 𝑓 defined in
(5.7) is called the gravitational potential.
∫ 2𝜋
= (cos2 𝑡 + sin2 𝑡)d𝑡
0
= 2𝜋.
□
An interpretation of the flux of a vector field is provided by the following
situation in fluid dynamics: Let 𝑉 (𝑥, 𝑦) denote the velocity field of a plane fluid
in some region 𝑈 in ℝ2 containing the simple closed curve 𝐶. Then, at each
point (𝑥, 𝑦) in 𝑈 , 𝑉 (𝑥, 𝑦) gives the velocity of the fluid as it goes through that
point in units of length per unit time. Suppose we know the density of the fluid
as a function, 𝜌(𝑥, 𝑦), of the position of the fluid in 𝑈 (this is a scalar field) in
units of mass per unit area (since this is a two–dimensional fluid). Then, the
vector field
𝐹 (𝑥, 𝑦) = 𝜌(𝑥, 𝑦)𝑉 (𝑥, 𝑦),
in units of mass per unit length per unit time, gives the rate of fluid flow per
unit length at the point (𝑥, 𝑦). The integrand
𝐹 ⋅𝑛
ˆd𝑠,
in the flux definition in (5.8), is then in units of mass per unit time and measures
the amount of fluid that crosses a section of the curve 𝐶 of length d𝑠 in the
outward normal direction. The flux then gives the rate at which the fluid is
crossing the curve 𝐶 from the inside to the outside; in other words, the flux
gives the rate of flow of fluid out of the region bounded by 𝐶.
Denoting the vector 𝑤𝑝 by (𝐹1 (𝑝), 𝐹2 (𝑝), . . . , 𝐹𝑛 (𝑝)), we can then write the ex-
pression in (5.9) as
Combining the result in (5.12) in Example 5.5.3 with that of (5.13) in Ex-
ample 5.5.4, we see that for a smooth function 𝑓 : 𝑈 → ℝ,
∂𝑓 ∂𝑓 ∂𝑓
𝑑𝑓𝑝 (ℎ) = (𝑝) 𝑑𝑥1 (ℎ) + (𝑝) 𝑑𝑥2 (ℎ) + ⋅ ⋅ ⋅ + (𝑝) 𝑑𝑥𝑛 (ℎ),
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
∂𝑓 ∂𝑓 ∂𝑓
𝑑𝑓𝑝 = (𝑝) 𝑑𝑥1 + (𝑝) 𝑑𝑥2 + ⋅ ⋅ ⋅ + (𝑝) 𝑑𝑥𝑛 ,
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
for 𝑝 ∈ 𝑈 , or
∂𝑓 ∂𝑓 ∂𝑓
𝑑𝑓 = 𝑑𝑥1 + 𝑑𝑥2 + ⋅ ⋅ ⋅ + 𝑑𝑥𝑛 , (5.14)
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
which gives an interpretation of the differential of a smooth function, 𝑓 , as
a differential 1–form. The expression in (5.14) displays 𝑑𝑓 as a linear combi-
nation of the set of differential 1–forms {𝑑𝑥1 , 𝑑𝑥2 , . . . , 𝑑𝑥𝑛 }. In fact, the set
{𝑑𝑥1 , 𝑑𝑥2 , . . . , 𝑑𝑥𝑛 } is a basis for the space of differential 1–forms. Thus, any
differential 1–form, 𝜔, can be written as
Solution: We compute
∫
𝜔([𝑃1 , 𝑃2 ]) = 𝑦𝑧d𝑥 + 𝑥𝑧d𝑦 + 𝑥𝑦d𝑧,
[𝑃1 ,𝑃2 ]
for 0 ⩽ 𝑡 ⩽ 1. Then, ⎧
⎨ 𝑑𝑥 = 2 𝑑𝑡
𝑑𝑦 = 𝑑𝑡
𝑑𝑧 = 𝑑𝑡,
⎩
and
∫ ∫ 1
𝑦𝑧d𝑥 + 𝑥𝑧d𝑦 + 𝑥𝑦d𝑧 = [2(1 + 𝑡)𝑡 + (1 + 2𝑡)𝑡 + (1 + 2𝑡)(1 + 𝑡)]d𝑡
𝐶 0
∫ 1
= (2𝑡 + 2𝑡2 + 𝑡 + 2𝑡2 + 1 + 𝑡 + 2𝑡 + 2𝑡2 )d𝑡
0
∫ 1
= (1 + 6𝑡 + 6𝑡2 )d𝑡
0
= 6.
Example 5.5.7. Let 𝜔 = 𝑘1 𝑑𝑥1 +𝑘2 𝑑𝑥2 +⋅ ⋅ ⋅+𝑘𝑛 𝑑𝑥𝑛 , where 𝑘1 , 𝑘2 , . . . , 𝑘𝑛 are
real constants, be a constant differential 1–form. For any two distinct points,
𝑃𝑜 and 𝑃1 , in ℝ𝑛 , compute 𝜔([𝑃𝑜 , 𝑃1 ])
Compute
∫
𝜔([𝑃𝑜 , 𝑃1 ]) = 𝐹 ⋅ d→
−
𝑟
[𝑃𝑜 ,𝑃1 ]
∫ 1
= 𝐹 (𝜎(𝑡)) ⋅ 𝜎 ′ (𝑡) d𝑡,
0
5.5. DIFFERENTIAL FORMS 79
where
−−→
𝜎(𝑡) = 𝑂𝑃𝑜 + 𝑡𝑣, for 0 ⩽ 𝑡 ⩽ 1,
−−−→
where 𝑣 = 𝑃𝑜 𝑃1 , the vector that goes from 𝑃𝑜 to 𝑃1 . Thus,
𝜔([𝑃𝑜 , 𝑃1 ]) = 𝐾 ⋅ 𝑣,
∂𝑓 ∂𝑓 ∂𝑓
𝑑𝑓 = 𝑑𝑥1 + 𝑑𝑥2 + ⋅ ⋅ ⋅ + 𝑑𝑥𝑛 .
∂𝑥1 ∂𝑥2 ∂𝑥𝑛
∫ 1
= ∇𝑓 (𝜎(𝑡)) ⋅ 𝜎 ′ (𝑡) d𝑡,
0
where
−−→ −−−→
𝜎(𝑡) = 𝑂𝑃1 + 𝑡𝑃1 𝑃2 , 0 ⩽ 𝑡 ⩽ 1.
Thus, by the Chain Rule,
∫ ∫ 1
𝑑
𝑑𝑓 = [𝑓 (𝜎(𝑡))] d𝑡
[𝑃1 ,𝑃2 ] 0 𝑑𝑡
= 𝑓 (𝑃2 ) − 𝑓 (𝑃1 ).
𝐵(𝑣, 𝑤) = 𝐵(𝑎ˆ𝑖 + 𝑏ˆ
𝑗, 𝑐ˆ𝑖 + 𝑑ˆ
𝑗)
= 𝜆(𝑎𝑑 − 𝑏𝑐)
( )
𝑎 𝑐
= 𝜆 det .
𝑏 𝑑
5.5. DIFFERENTIAL FORMS 81
where [ 𝑣 𝑤 ] denotes the 2 × 2 matrix whose first column are the entries of 𝑣,
and whose second column are the entries of 𝑤.
Example 5.5.17 (Skew–Symmetric Forms in ℝ3 ). Let 𝐵 : ℝ3 × ℝ3 → ℝ be a
skew–symmetric bilinear form on ℝ3 . We than have that Then
and
𝐵(ˆ
𝑗, ˆ𝑖) = −𝐵(ˆ𝑖, ˆ
𝑗),
𝐵(ˆ
𝑘, ˆ𝑖) = −𝐵(ˆ𝑖, ˆ
𝑘), (5.18)
𝐵(ˆ
𝑘, ˆ
𝑗) = −𝐵(ˆ
𝑗, ˆ
𝑘).
Set
𝜆1 = 𝐵(ˆ
𝑗, ˆ
𝑘),
𝜆2 = 𝐵(ˆ𝑖, ˆ
𝑘), (5.19)
𝜆3 = 𝐵(ˆ𝑖, ˆ
𝑗).
Then, for any vectors 𝑣 = 𝑎1ˆ𝑖 + 𝑎2ˆ
𝑗 + 𝑎3 ˆ
𝑘 and 𝑤 = 𝑏1ˆ𝑖 + 𝑏2ˆ 𝑘 in ℝ3 , we
𝑗 + 𝑏3 ˆ
have that
𝐵(𝑣, 𝑤) = 𝐵(𝑎1ˆ𝑖 + 𝑎2ˆ
𝑗 + 𝑎3 ˆ
𝑘, 𝑏1ˆ𝑖 + 𝑏2ˆ
𝑗 + 𝑏3 ˆ
𝑘)
= 𝑎1 𝑏2 𝐵(ˆ𝑖, ˆ
𝑗) + 𝑎1 𝑏3 𝐵(ˆ𝑖, ˆ𝑘)
+𝑎2 𝑏1 𝐵(ˆ 𝑗, ˆ𝑖) + 𝑎2 𝑏3 𝐵(ˆ𝑗, ˆ
𝑘)
+𝑎3 𝑏1 𝐵(𝑘, 𝑖) + 𝑎3 𝑏2 𝐵(ˆ
ˆ ˆ 𝑘, ˆ
𝑗),
𝐵(𝑣, 𝑤) = 𝑎2 𝑏3 𝐵(ˆ
𝑗, ˆ
𝑘) + 𝑎3 𝑏2 𝐵(ˆ𝑘, ˆ
𝑗)
+𝑎3 𝑏1 𝐵(𝑘, 𝑖) + 𝑎1 𝑏3 𝐵(ˆ𝑖, ˆ
ˆ ˆ 𝑘) (5.20)
+𝑎1 𝑏2 𝐵(ˆ𝑖, ˆ
𝑗) + 𝑎2 𝑏1 𝐵(ˆ
𝑗, ˆ𝑖).
Recognizing the term on the right–hand side of (5.21) as the triple scalar product
of the vector Λ = 𝜆1ˆ𝑖 + 𝜆2ˆ
𝑗 + 𝜆3 ˆ
𝑘 and the vectors 𝑣 and 𝑤, we see that we have
shown that for every skew–symmetric, bilinear form, 𝐵 : ℝ3 × ℝ3 → ℝ, there
exists a vector Λ ∈ ℝ3 such that
𝐵(𝑣, 𝑤) = Λ ⋅ (𝑣 × 𝑤), for all 𝑣, 𝑤 ∈ ℝ3 . (5.22)
so that
(𝜔 ∧ 𝜂)𝑝 (𝑐1 𝑣1 + 𝑐2 𝑣2 , 𝑤) = 𝑐1 [𝜔𝑝 (𝑣1 )𝜂𝑝 (𝑤) − 𝜔𝑝 (𝑤)𝜂𝑝 (𝑣1 )]
+𝑐2 [𝜔𝑝 (𝑣2 )𝜂𝑝 (𝑤) − 𝜔𝑝 (𝑤)𝜂𝑝 (𝑣2 )]
and
−−−→
𝑤 = 𝑃𝑜 𝑃2 = (𝑥2 − 𝑥𝑜 )ˆ𝑖 + (𝑦2 − 𝑦𝑜 )ˆ
𝑗.
Then, according to (5.25) in Definition 5.5.21,
where we have used the result of Example 5.5.11. We then have that
𝑥1 − 𝑥𝑜 𝑥2 − 𝑥𝑜
(𝑑𝑥 ∧ 𝑑𝑦)(𝑣, 𝑤) = ,
𝑦1 − 𝑦𝑜 𝑦2 − 𝑦𝑜
84 CHAPTER 5. INTEGRATION
which is the determinant of the 2 × 2 matrix, [𝑣 𝑤], whose columns are the
vectors 𝑣 and 𝑤. In other words,
We have therefore shown that the (𝑑𝑥 ∧ 𝑑𝑦)(𝑣, 𝑤) gives the signed area of the
parallelogram determined by the vectors 𝑣 and 𝑤.
and
−−−→
𝑤 = 𝑃𝑜 𝑃2 = (𝑥2 − 𝑥𝑜 )ˆ𝑖 + (𝑦2 − 𝑦𝑜 )ˆ
𝑗 + (𝑧1 − 𝑧𝑜 )ˆ
𝑘.
Then, as in Example 5.5.23, we compute
𝑥1 − 𝑥𝑜 𝑥2 − 𝑥𝑜
(𝑑𝑥 ∧ 𝑑𝑦)(𝑣, 𝑤) = ,
𝑦1 − 𝑦𝑜 𝑦2 − 𝑦𝑜
Similarly, we compute
𝑦1 − 𝑦𝑜 𝑧1 − 𝑧𝑜
(𝑑𝑦 ∧ 𝑑𝑧)(𝑣, 𝑤) = , (5.28)
𝑦2 − 𝑦𝑜 𝑧2 − 𝑧𝑜
and
𝑧1 − 𝑧𝑜 𝑥1 − 𝑥𝑜
(𝑑𝑧 ∧ 𝑑𝑥)(𝑣, 𝑤) = ,
𝑧2 − 𝑧𝑜 𝑥2 − 𝑥𝑜
or
𝑥1 − 𝑥𝑜 𝑧1 − 𝑧𝑜
(𝑑𝑧 ∧ 𝑑𝑥)(𝑣, 𝑤) = − . (5.29)
𝑥2 − 𝑥𝑜 𝑧2 − 𝑧𝑜
We recognize in (5.28), (5.29) and (5.27) the components of the cross product
of the vectors 𝑣 and 𝑤,
𝑦1 − 𝑦𝑜 𝑧1 − 𝑧𝑜 𝑥1 − 𝑥𝑜 𝑧1 − 𝑧𝑜 𝑥1 − 𝑥𝑜 𝑦1 − 𝑦𝑜
𝑣×𝑤 = ˆ𝑖− 𝑗+
ˆ 𝑘.
ˆ
𝑦2 − 𝑦𝑜 𝑧2 − 𝑧𝑜 𝑥2 − 𝑥𝑜 𝑧2 − 𝑧𝑜 𝑥2 − 𝑥𝑜 𝑦2 − 𝑦𝑜
5.5. DIFFERENTIAL FORMS 85
(𝑑𝑧 ∧ 𝑑𝑥)(𝑣, 𝑤) = (𝑣 × 𝑤) ⋅ ˆ
𝑗, (5.31)
and
(𝑑𝑥 ∧ 𝑑𝑦)(𝑣, 𝑤) = (𝑣 × 𝑤) ⋅ ˆ
𝑘. (5.32)
d𝑥 ∧ d𝑦(𝑇 ) = ± area(𝑇 ).
We denote this by ∫
d𝑥 ∧ d𝑦 = signed area of 𝑇. (5.33)
𝑇
Example 5.5.27. Let 𝑃𝑜 (0, 0), 𝑃1 (1, 2) and 𝑃2 (2, 1) and let 𝑇 ∫
= [𝑃𝑜 , 𝑃1 , 𝑃2 ]
denote the oriented triangle generated by those points. Evaluate d𝑥 ∧ d𝑦.
𝑇
86 CHAPTER 5. INTEGRATION
Solution: Set
⎛⎞ ⎛⎞
4 5
−−−→ −−−→
𝑣 = 𝑃𝑜 𝑃1 = ⎝ 1 ⎠ and 𝑤 = 𝑃𝑜 𝑃2 = ⎝ 6 ⎠ ,
−1 −1
and compute
ˆ𝑖 𝑗
ˆ 𝑘
ˆ
𝑣×𝑤 = 4 1 −1 = (−2+6) ˆ𝑖−(−8+5) ˆ
𝑗+(24−5) ˆ
𝑘 = 4 ˆ𝑖+3 ˆ
𝑗+19 ˆ
𝑘.
5 6 −2
It then follows that ∫
d𝑦 ∧ d𝑧 = 2,
𝑇
∫
3
d𝑧 ∧ d𝑥 =
𝑇 2
and ∫
19
d𝑥 ∧ d𝑦 = .
𝑇 2
□
𝜔𝑝 (𝑣, 𝑤) = 𝜔𝑝 (𝑎1 ˆ𝑖 + 𝑎2 ˆ
𝑗 + 𝑎3 , 𝑏1 ˆ𝑖 + 𝑏2 ˆ
𝑗 + 𝑏3 ˆ
𝑘)
= 𝑎1 𝑏2 𝜔𝑝 (ˆ𝑖, ˆ
𝑗) + 𝑎1 𝑏3 𝜔𝑝 (ˆ𝑖, ˆ
𝑘) (5.37)
𝑎2 𝑏1 𝜔𝑝 (ˆ𝑗, ˆ𝑖) + 𝑎2 𝑏3 𝜔𝑝 (ˆ
𝑗, ˆ
𝑘)+
𝑎3 𝑏1 𝜔𝑝 (𝑘, 𝑖) + 𝑎3 𝑏2 𝜔𝑝 (ˆ
ˆ ˆ 𝑘, ˆ
𝑗),
88 CHAPTER 5. INTEGRATION
𝜔𝑝 (ˆ𝑖, ˆ𝑖) = 𝜔𝑝 (ˆ
𝑗, ˆ
𝑗) = 𝜔𝑝 (ˆ
𝑘, ˆ
𝑘) = 0,
𝜔𝑝 (𝑣, 𝑤) = 𝑎2 𝑏3 𝜔𝑝 (ˆ
𝑗, ˆ
𝑘) + 𝑎3 𝑏2 𝜔𝑝 (ˆ
𝑘, ˆ
𝑗)
+𝑎1 𝑏3 𝜔𝑝 (ˆ𝑖, ˆ
𝑘) + 𝑎3 𝑏1 𝜔𝑝 (ˆ
𝑘, ˆ𝑖)
+𝑎1 𝑏2 𝜔𝑝 (𝑖, 𝑗) + 𝑎2 𝑏1 𝜔𝑝 (ˆ
ˆ ˆ 𝑗, ˆ𝑖)
(5.38)
= 𝜔𝑝 (ˆ 𝑘)(𝑎2 𝑏3 − 𝑎3 𝑏2 )
𝑗, ˆ
+𝜔𝑝 (ˆ 𝑘, ˆ𝑖)(𝑎3 𝑏1 − 𝑎1 𝑏3 )
𝑗)(𝑎1 𝑏2 − 𝑎2 𝑏1 ).
+𝜔𝑝 (ˆ𝑖, ˆ
𝜔𝑝 (𝑣, 𝑤) = 𝜔𝑝 (ˆ 𝑘) 𝑑𝑦 ∧ 𝑑𝑧(𝑣, 𝑤)
𝑗, ˆ
+𝜔𝑝 (ˆ 𝑘, ˆ𝑖) 𝑑𝑧 ∧ 𝑑𝑥(𝑣, 𝑤)
𝑗) 𝑑𝑥 ∧ 𝑑𝑦(𝑣, 𝑤),
+𝜔𝑝 (ˆ𝑖, ˆ
𝜔𝑝 = 𝜔𝑝 (ˆ 𝑘) 𝑑𝑦 ∧ 𝑑𝑧 + 𝜔𝑝 (ˆ
𝑗, ˆ 𝑘, ˆ𝑖) 𝑑𝑧 ∧ 𝑑𝑥 + 𝜔𝑝 (ˆ𝑖, ˆ
𝑗) 𝑑𝑥 ∧ 𝑑𝑦. (5.42)
Setting
𝐹1 (𝑝) = 𝜔𝑝 (ˆ
𝑗, ˆ
𝑘),
𝐹2 (𝑝) = 𝜔𝑝 (ˆ
𝑘, ˆ𝑖),
𝐹3 (𝑝) = 𝜔𝑝 (ˆ𝑖, ˆ
𝑗),
5.6. CALCULUS OF DIFFERENTIAL FORMS 89
which shows that every differential 2–form in ℝ3 is in the span of the set in
(5.36).
To show that the representation in (5.43) is unique, assume that
the differential 2–form that maps every pair of vectors (𝑣, 𝑤) ∈ ℝ3 × ℝ3 to the
real number 0. Then, applying the form in (5.44) to the pair (ˆ 𝑗, ˆ
𝑘) we obtain
that
respectively. Thus, the set in (5.36) is also linearly independent; hence, the
representation in (5.43) is unique.
Consequently,
𝑑𝑦 ∧ 𝑑𝑥 = −𝑑𝑥 ∧ 𝑑𝑦. (5.45)
From this we can deduce that
𝑑𝑥 ∧ 𝑑𝑥 = 0. (5.46)
90 CHAPTER 5. INTEGRATION
𝑃 𝑑𝑥 + 𝑄 𝑑𝑦, (5.47)
where 𝑃 and 𝑄 are smooth scalar fields. We can also multiply the differential
1–from in (5.47) by the 1–form 𝑑𝑥:
(𝑃 𝑑𝑥 + 𝑄 𝑑𝑦) ∧ 𝑑𝑥 = 𝑃 𝑑𝑥 ∧ 𝑑𝑥 + 𝑄 𝑑𝑦 ∧ d𝑥 = −𝑄 𝑑𝑥 ∧ 𝑑𝑦,
𝜔 = 𝑃 (𝑥, 𝑦) 𝑑𝑥 + 𝑄(𝑥, 𝑦) 𝑑𝑦
d𝜔 = (𝑑𝑃 ) ∧ 𝑑𝑥 + (𝑑𝑄) ∧ 𝑑𝑦
( ) ( )
∂𝑃 ∂𝑃 ∂𝑄 ∂𝑄
= 𝑑𝑥 + 𝑑𝑦 ∧ 𝑑𝑥 + 𝑑𝑥 + ∧ d𝑦 𝑑𝑦
∂𝑥 ∂𝑦 ∂𝑥 ∂𝑦
∂𝑃 ∂𝑃 ∂𝑄 ∂𝑄
= 𝑑𝑥 ∧ 𝑑𝑥 + 𝑑𝑦 ∧ 𝑑𝑥 + 𝑑𝑥 ∧ 𝑑𝑦 + 𝑑𝑦 ∧ 𝑑𝑦
∂𝑥 ∂𝑦 ∂𝑥 ∂𝑦
( )
∂𝑄 ∂𝑃
= − 𝑑𝑥 ∧ 𝑑𝑦,
∂𝑥 ∂𝑦
where we have used (5.45) and (5.46). Thus, the differential of the 1–form
𝜔 = 𝑃 𝑑𝑥 + 𝑄 𝑑𝑦
𝑇 = [𝑃1 , 𝑃2 , 𝑃3 ],
For the case in which 𝑇 has a positive orientation, we will denote the value of
∫
𝑓 (𝑥, 𝑦)d𝑥 ∧ d𝑦 by
𝑇 ∫
𝑓 (𝑥, 𝑦)d𝑥d𝑦 (5.49)
𝑇
and call it the double integral of 𝑓 over 𝑇 . In this sense, we then have that
∫ ∫
𝑓 (𝑥, 𝑦)d𝑦 ∧ d𝑥 = − 𝑓 (𝑥, 𝑦)d𝑥d𝑦,
𝑇 𝑇
(0, 1)
@
@ 𝑥+𝑦 =1
@
@
@
@
(0, 0) (1, 0) 𝑥
yields a function of 𝑥 for 𝑥 ∈ [0, 1]; call this function 𝑔; that is,
∫ 1−𝑥
𝑔(𝑥) = 𝑓 (𝑥, 𝑦) d𝑦 for all 𝑥 ∈ [0, 1];
0
Then, ∫ ∫ 1
𝑓 (𝑥, 𝑦)d𝑥d𝑦 = 𝑔(𝑥) d𝑥.
𝑈 0
We could also do the integration with respect to 𝑥 first, then integrate with
respect to 𝑦:
∫ ∫ 1 {∫ 1−𝑦 }
𝑓 (𝑥, 𝑦)d𝑥d𝑦 = 𝑓 (𝑥, 𝑦) d𝑥 d𝑦. (5.51)
𝑈 0 0
In this case the inner integral yields a function of 𝑦 which can then be integrated
from 0 to 1.
Observe that the iterated integrals in (5.50) and (5.51) correspond to alter-
nate descriptions of 𝑈 as
𝑈 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 0 ⩽ 𝑥 ⩽ 1, 0 ⩽ 𝑦 ⩽ 1 − 𝑥}
or
𝑈 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 0 ⩽ 𝑥 ⩽ 1 − 𝑦, 0 ⩽ 𝑦 ⩽ 1},
respectively.
The fact that the iterated integrals in equations (5.50) and (5.51) yield the
same value, at least for the case in which 𝑓 is continuous on a region containing
𝑈 , is a special case of a theorem in Advanced Calculus or Real Analysis known
as Fubini’s Theorem.
5.7. EVALUATING 2–FORMS: DOUBLE INTEGRALS 93
∫
Example 5.7.1. Evaluate 𝑥 d𝑥d𝑦.
𝑈
∫ 1 [ ]1−𝑥
= 𝑥𝑦 d𝑥
0 0
∫ 1
= 𝑥(1 − 𝑥) d𝑥
0
∫ 1
= (𝑥 − 𝑥2 ) d𝑥
0
1
= .
6
We could have also used the iterated integral in (5.51):
∫ ∫ 1 {∫ 1−𝑦 }
𝑥 d𝑥d𝑦 = 𝑥 d𝑥 d𝑦
𝑈 0 0
∫ 1 [1 ]1−𝑦
= 𝑥2 d𝑦
0 2 0
∫ 1
1
= (1 − 𝑦)2 d𝑦
2 0
∫ 0
1
= − 𝑢2 d𝑥
2 1
∫ 1
1
= 𝑢2 d𝑢
2 0
1
= .
6
□
Iterated integrals can be used to evaluate double–integrals over plane regions
other than triangles. For instance, suppose a region, 𝑅, is bounded by the
vertical lines 𝑥 = 𝑎 and 𝑥 = 𝑏, where 𝑎 < 𝑏, and by the graphs of two functions
𝑔1 (𝑥) and 𝑔2 (𝑥), where 𝑔1 (𝑥) ⩽ 𝑔2 (𝑥) for 𝑎 ⩽ 𝑥 ⩽ 𝑏; that is
then, {∫ }
∫ ∫ 𝑏 𝑔2 (𝑥)
𝑓 (𝑥, 𝑦) d𝑥d𝑦 = 𝑓 (𝑥, 𝑦) d𝑦 d𝑥.
𝑅 𝑎 𝑔1 (𝑥)
Example 5.7.2. Let 𝑅 denote the region in the first quadrant bounded ∫ by the
2 2
unit circle, 𝑥 + 𝑦 = 1; that is, 𝑅 is the quarter unit disc. Evaluate 𝑦 d𝑥d𝑦.
𝑅
so that √
∫ ∫ 1 ∫ 1−𝑥2
𝑦 d𝑥d𝑦 = 𝑦 d𝑦d𝑥
𝑅 0 0
√
∫ 1 1−𝑥2
1 2
= 𝑦 d𝑥
0 2 0
∫ 1
1
= (1 − 𝑥2 ) d𝑥
2 0
1
=
3
□
Example 5.7.3. Identify the region, 𝑅, in the plane in which the following
iterated integral ∫ 1∫ 1
1
√ d𝑥d𝑦
0 𝑦 1 + 𝑥2
is computed. Change the order of integration and then evaluate the double inte-
gral ∫
1
√ d𝑥d𝑦.
𝑅 1 + 𝑥2
Solution: In this case, the region 𝑅 is
𝑅 = {(𝑥, 𝑦) ∈ ℝ2 ∣ 𝑦 ⩽ 𝑥 ⩽ 1, 1 ⩽ 𝑦 ⩽ 1}.
5.7. EVALUATING 2–FORMS: DOUBLE INTEGRALS 95
𝑥=𝑦 𝑥=1
∫ 1 𝑥
1
= √ 𝑦 d𝑥
0 1 + 𝑥2 0
∫ 1
1
= √ 𝑥 d𝑥
0 1 + 𝑥2
∫ 1
1
= √ 2𝑥 d𝑥
0 2 1 + 𝑥2
∫ 2
1
= √ d𝑢
1 2 𝑢
√ 2
= 𝑢
1
√
= 2 − 1.
□
2
If 𝑅 is a bounded region of ℝ , and 𝑓 (𝑥, 𝑦) ⩾ 0 for all (𝑥, 𝑦) ∈ 𝑅, then
∫
𝑓 (𝑥, 𝑦) d𝑥d𝑦
𝑅
gives the volume of the three dimensional solid that lies below the graph of the
surface 𝑧 = 𝑓 (𝑥, 𝑦) and above the region 𝑅.
96 CHAPTER 5. INTEGRATION
Example 5.7.4. Let 𝑎, 𝑏 and 𝑐 be positive real numbers. Compute the volume
of the tetrahedron whose base is the triangle 𝑇 = [(0, 0), (𝑎, 0), (0, 𝑏)] and which
lies below the plane
𝑥 𝑦 𝑧
+ + = 1.
𝑎 𝑏 𝑐
∫
Solution: We need to evaluate 𝑧 d𝑥d𝑦, where
𝑇
( 𝑥 𝑦)
𝑧 =𝑐 1− − .
𝑎 𝑏
Then,
∫ ∫ (
𝑥 𝑦)
𝑧 d𝑥d𝑦 = 𝑐 1− − d𝑥d𝑦
𝑇 𝑇 𝑎 𝑏
∫ 𝑎 ∫ 𝑏(1−𝑥/𝑎) ( 𝑥 𝑦)
= 𝑐 1− − d𝑦d𝑥
0 0 𝑎 𝑏
𝑎 ]𝑏(1−𝑥/𝑎)
𝑥𝑦 𝑦 2
∫ [
= 𝑐 𝑦− − d𝑥
0 𝑎 2𝑏 0
∫ 𝑎 [ ( ]
𝑥) 𝑥 ( 𝑥) 1 2( 𝑥 )2
= 𝑐 𝑏 1− − 𝑏 1− − 𝑏 1− d𝑥
0 𝑎 𝑎 𝑎 2𝑏 𝑎
𝑎
𝑥2
∫ ( )
1 𝑥
= 𝑏𝑐 − + 2 d𝑥
0 2 𝑎 2𝑎
[𝑎 𝑎 𝑎]
= 𝑏𝑐 − +
2 2 6
𝑎𝑏𝑐
= .
6
□
Proof of Green’s Theorem for the Unit Triangle in ℝ2 . We shall first prove Propo-
sition 5.8.1 for the unit triangle 𝑈 = [(0, 0), (1, 0), (0, 1)] = [𝑃1 , 𝑃2 , 𝑃3 ]:
∫ ( ) ∫
∂𝑄 ∂𝑃
− d𝑥d𝑦 = 𝑃 d𝑥 + 𝑄d𝑦, (5.54)
𝑈 ∂𝑥 ∂𝑦 ∂𝑈
where 𝑃 and 𝑄 are 𝐶 1 scalar fields defined on some region containing 𝑈 , and ∂𝑈
is made up of the directed line segments [𝑃1 , 𝑃2 ], [𝑃2 , 𝑃3 ] and [𝑃3 , 𝑃1 ] traversed
in the counterclockwise sense.
We will prove separately that
∫ ∫
∂𝑄
d𝑥d𝑦 = 𝑄d𝑦, (5.55)
𝑈 ∂𝑥 ∂𝑈
and ∫ ∫
∂𝑃
− d𝑥d𝑦 = 𝑃 d𝑥. (5.56)
𝑈 ∂𝑦 ∂𝑈
𝑃3 @
@ 𝑥+𝑦 =1
@
@
@
@
𝑃1 𝑃2 𝑥
or ∫ ∫ ∫
𝑄d𝑦 = 𝑄d𝑦 + 𝑄d𝑦, (5.58)
∂𝑈 [𝑃2 ,𝑃3 ] [𝑃3 ,𝑃1 ]
since d𝑦 = 0 on [𝑃1 , 𝑃2 ].
Now, parametrize [𝑃2 , 𝑃3 ] by
{
𝑥=1−𝑦
𝑦 = 𝑦,
Parametrizing [𝑃3 , 𝑃1 ] by
{
𝑥=0
𝑦 = 1 − 𝑡,
Comparing the left–hand sides on the equations (5.61) and (5.57), we see that
(5.55) is true. A similar calculation shows that (5.56) is also true. Hence,
Proposition 5.8.1 is proved for the unit triangle 𝑈 .
5.9. CHANGING VARIABLES 99
𝑏 H
HH
HH
Δ𝑥 H
Δ𝑦 HH
(𝑥, 𝑦) H
H H
H
𝑎 𝑥
(0, 1)
@
@
@
Δ𝑢 @
Δ𝑣@
(𝑢, 𝑣) @
(0, 0) (1, 0) 𝑢
for 0 ⩽ 𝑡 ⩽ 1, which is a parametrization of the line segment [(0, 0), (𝑎, 0)] in
the 𝑥𝑦–plane.
Similarly, the line segment [(1, 0), (0, 1)] in the 𝑢𝑣–plane, parametrized by
{
𝑢=1−𝑡
𝑣 = 𝑡,
for 0 ⩽ 𝑡 ⩽ 1, which is a parametrization of the line segment [(𝑎, 0), (0, 𝑏)] in
the 𝑥𝑦–plane.
Similar considerations show that [(0, 1), (0, 0)] gets mapped to [(0, 𝑏), (0, 0)]
under the action of Φ on ℝ2 .
Writing ( ) ( ) ( )
𝑥(𝑢, 𝑣) 𝑢 𝑢
=Φ for all ∈ ℝ2 ,
𝑦(𝑢, 𝑣) 𝑣 𝑣
we can express the integrand in the double integral in (5.62) as a function of 𝑢
and 𝑣:
𝑓 (𝑥(𝑢, 𝑣), 𝑦(𝑢, 𝑣)) for (𝑢, 𝑣) in 𝑈.
We presently see how the differential 2–form d𝑥d𝑦 can be expressed in terms of
d𝑢d𝑣. To do this consider the small rectangle of area Δ𝑢Δ𝑣 and lower left–hand
corner at (𝑢, 𝑣) pictured in Figure 5.9.9. We see where the vector field Φ maps
this rectangle in the 𝑥𝑦–plane. In this case, it happens to be a rectangle with
5.9. CHANGING VARIABLES 101
lower–left hand corner Φ(𝑢, 𝑣) = (𝑥, 𝑦) and dimensions 𝑎Δ𝑢 × 𝑏Δ𝑣. In general,
however, the image of the Δ𝑢 × Δ𝑣 rectangle under a change of coordinates Φ
will be a plane region bounded by curves like the one pictured in Figure 5.9.10.
In the general case, we approximate the area of the image region by the area
𝑏 H
H
HH
H
HH
H
HH
(𝑥, 𝑦) H
H
𝑎 𝑥
of the parallelogram spanned by vectors tangent to the image curves of the line
segments [(𝑢, 𝑣), (𝑢 + Δ𝑢, 𝑣)] and [(𝑢, 𝑣), (𝑢, 𝑣 + Δ𝑣)] under the map Φ at the
point (𝑢, 𝑣). The curves are given parametrically by
and
𝛾(𝑣) = Φ(𝑢, 𝑣) = (𝑥(𝑢, 𝑣), 𝑦(𝑢, 𝑣)) for 𝑣 ⩽ 𝑣 ⩽ 𝑣 + Δ𝑣.
and
( )
∂𝑥 ˆ ∂𝑦 ˆ
Δ𝑣 𝛾 ′ (𝑣) = Δ𝑣 𝑖+ 𝑗 ,
∂𝑣 ∂𝑣
where we have scaled by Δ𝑢 and Δ𝑣, respectively, by virtue of the linear ap-
proximation provided by the derivative maps 𝐷𝜎(𝑢) and 𝐷𝛾(𝑣), respectively.
The area of the image rectangle can then be approximated by the norm of the
cross product of the tangent vectors:
= ∥𝜎 ′ (𝑢) × 𝛾 ′ (𝑣)∥Δ𝑢Δ𝑣
102 CHAPTER 5. INTEGRATION
∂𝑥 ∂𝑦 ˆ ˆ ∂𝑦 ∂𝑥 ˆ ˆ
= 𝑖×𝑗+ 𝑗×𝑖
∂𝑢 ∂𝑣 ∂𝑢 ∂𝑣
( )
∂𝑥 ∂𝑦 ∂𝑦 ∂𝑥 ˆ
= − 𝑘
∂𝑢 ∂𝑣 ∂𝑢 ∂𝑣
∂(𝑥, 𝑦) ˆ
= 𝑘,
∂(𝑢, 𝑣)
∂(𝑥, 𝑦)
where denotes the determinant of the Jacobian matrix of the Φ at (𝑢, 𝑣).
∂(𝑢, 𝑣)
It then follows that
∂(𝑥, 𝑦)
Δ𝑥Δ𝑦 ≈ Δ𝑢Δ𝑣,
∂(𝑢, 𝑣)
which translates in terms of differential forms to
∂(𝑥, 𝑦)
d𝑥d𝑦 = d𝑢d𝑣.
∂(𝑢, 𝑣)
We therefore obtain the Change of Variables Formula
∫ ∫
∂(𝑥, 𝑦)
𝑓 (𝑥, 𝑦) d𝑥d𝑦 = 𝑓 (𝑥(𝑢, 𝑣), 𝑦(𝑢, 𝑣)) d𝑢d𝑣. (5.63)
𝑇 𝑈 ∂(𝑢, 𝑣)
This formula works for any regions 𝑅 and 𝐷 in the plane for which there is a
change of coordinates Φ : ℝ2 → ℝ2 such that Φ(𝐷) = 𝑅:
∫ ∫
∂(𝑥, 𝑦)
𝑓 (𝑥, 𝑦) d𝑥d𝑦 = 𝑓 (𝑥(𝑢, 𝑣), 𝑦(𝑢, 𝑣)) d𝑢d𝑣. (5.64)
𝑅 𝐷 ∂(𝑢, 𝑣)
Example 5.9.1. For the case in which 𝑇 = [(0, 0), (𝑎, 0), (0, 𝑏)] and 𝑈 is the
unit triangle in ℝ2 , and Φ is given by
( ) ( ) ( )
𝑢 𝑎𝑢 𝑢
Φ = for all ∈ ℝ2 ,
𝑣 𝑏𝑣 𝑣
= 𝑟.
Hence,
∫ ∫
−𝑥2 −𝑦 2 2
𝑒 d𝑥d𝑦 = 𝑒−𝑟 𝑟 d𝑟d𝜃
𝑅 𝐷
∫ 2𝜋 ∫ 1
2
= 𝑒−𝑟 𝑟 d𝑟d 𝜃
0 0
∫ 2𝜋 [ ]1
1 2
= − 𝑒−𝑟 d𝜃
0 2 0
∫ 2𝜋
1
1 − 𝑒−1
( )
= d𝜃
2 0
= 𝜋 1 − 𝑒−1 .
( )
{𝑥 + 𝑡(𝑦 − 𝑥) ∈ ℝ𝑛 ∣ 0 ≤ 𝑡 ⩽ 1} ⊆ 𝐴
Example A.0.5. Prove that the ball 𝐵𝑟 (𝑂) = {𝑥 ∈ ℝ𝑛 ∣ ∥𝑥∥ < 𝑟} is a convex
subset of ℝ𝑛 .
Solution: Let 𝑥 and 𝑦 be in 𝐵𝑟 (𝑂); then, ∥𝑥∥ < 𝑟 and ∥𝑦∥ < 𝑟.
For 0 ⩽ 𝑡 ⩽ 1, consider
Thus, 𝑥 + 𝑡(𝑦 − 𝑥) ∈ 𝐵𝑟 (𝑂) for any 𝑡 ∈ [0, 1]. Since this is true for
any 𝑥, 𝑦 ∈ 𝐵𝑟 (𝑂), it follows that 𝐵𝑟 (𝑂) is convex. □
105
106 APPENDIX A. MEAN VALUE THEOREM
𝑦−𝑥
= ∇𝑓 (𝑧) ⋅ ∥𝑦 − 𝑥∥
∥𝑦 − 𝑥∥
= ∇𝑓 (𝑧) ⋅ 𝑢
ˆ ∥𝑦 − 𝑥∥
Reparametrizations
Theorem B.0.7. Let 𝐼 and 𝐽 denote open intervals of real numbers containing
closed and bounded intervals [𝑎, 𝑏] and [𝑐, 𝑑], respectively, and 𝛾1 : 𝐼 → ℝ𝑛 and
𝛾2 : 𝐽 → ℝ𝑛 be 𝐶 1 paths. Suppose that 𝐶 = 𝛾1 ([𝑎, 𝑏]) = 𝛾2 ([𝑐, 𝑑]) is a 𝐶 1 simple
curve parametrized by 𝛾1 and 𝛾2 . Then, there exists a differentiable function,
𝜏 : 𝐽 → 𝐼, such that
Remark B.0.9. Observe that the set 𝑝 + 𝑇𝑝 (𝐶) is the tangent line to the curve
𝐶 at 𝑝, hence the name “tangent space” for 𝑇𝑝 (𝐶).
107
108 APPENDIX B. REPARAMETRIZATIONS
𝑓 , which is defined on an open region containing 𝐶, the Chain Rule implies that
𝑓 ∘ 𝜎 is differentiable at 0 and
defines a linear map on the tangent space of 𝐶 at 𝑝. We will denote this linear
map by 𝑑𝑓𝑝 ; that is, 𝑑𝑓𝑝 : 𝑇𝑝 (𝐶) → ℝ is given by
Observe that we can then write, for ℎ ∈ ℝ with ∣ℎ∣ sufficiently small,
where
∣𝐸0 (ℎ)∣
lim = 0.
ℎ→0 ∣ℎ∣
Definition B.0.10. Let 𝐶 denote a 𝐶 1 curve parametrized by a 𝐶 1 path,
𝜎 : 𝐼 → ℝ𝑛 , where 𝐽 is an open interval containing 0 and such that 𝜎(0) = 𝑝 ∈ 𝐶.
We say that the function 𝑔 : 𝐶 → ℝ is differentiable at 𝑝 if there exists a linear
function
𝑑𝑔𝑝 : 𝑇𝑝 (𝐶) → ℝ
such that
(𝑔 ∘ 𝜎)(ℎ) = 𝑔(𝑝) + 𝑑𝑔𝑝 (ℎ𝜎 ′ (0)) + 𝐸𝑝 (ℎ),
where
∣𝐸𝑝 (ℎ)∣
lim = 0.
ℎ→0 ∣ℎ∣
We see from Definition B.0.10 that, if 𝑔 : 𝐶 → ℝ is differentiable at 𝑝, then
𝑔(𝜎(ℎ)) − 𝑔(𝑝)
lim
ℎ→0 ℎ
exists and equals 𝑑𝑔𝑝 (𝜎 ′ (0)). We have already seen that if 𝑓 is a 𝐶 1 scalar field
defined in an open region containing 𝐶, then
interior with 𝛾(0) = 𝑝. Here we are assuming that 𝛾 is one–to–one and onto 𝐶,
so that
𝑔 = 𝛾 −1 : 𝐶 → 𝐽
is defined. We claim that, since 𝛾 ′ (0) ∕= 0, according to the definition of 𝐶 1
parametrization in Definition 5.1.1 on page 61 in these notes, the function 𝑔 is
differentiable at 𝑝 according to Definition B.0.10. In order to prove this, we
first show that 𝑔 is continuous at 𝑝; that is,
Lemma B.0.11. Let 𝐶 be a 𝐶 1 curve parametrized by a 𝐶 1 map, 𝜎 : 𝐼 → ℝ𝑛 ,
where 𝐼 is an interval of real numbers containing 0 in its interior with 𝜎(0) = 𝑝.
Let 𝛾 : 𝐽 → ℝ𝑛 denote another 𝐶 1 parametrization of 𝐶, where 𝐽 is an interval
of real numbers containing 0 in its interior with 𝛾(0) = 𝑝. For every 𝑞 ∈ 𝐶,
define 𝑔(𝑞) = 𝜏 if and only if 𝛾(𝜏 ) = 𝑞. Then,
Proof: Write
𝜏 (ℎ) = 𝑔(𝜎(ℎ)), for ℎ ∈ 𝐼. (B.2)
We will show that
lim 𝜏 (ℎ) = 0; (B.3)
ℎ→0
𝑦𝑗′ (0) ∕= 0.
Next, use the differentiability of the function 𝑦𝑗 : 𝐽 → ℝ and the mean value
theorem to obtain 𝜃 ∈ (0, 1) such that
∣𝜏 (ℎ)∣ ⩽ 𝛿,
where
𝐸𝑗 (ℎ)
lim = 0. (B.16)
ℎ→0 ℎ
Consequently, using (B.13) and (B.15), if 𝜏 (ℎ)∣ ⩽ 𝛿,
The statement in (B.3) now follows from (B.17) and (B.16), since 𝑚 > 0 by
virtue of (B.14).
Lemma B.0.12. Let 𝐶, 𝜎 : 𝐼 → ℝ𝑛 and 𝛾 : 𝐽 → ℝ𝑛 be as in Lemma B.0.11. For
every 𝑞 ∈ 𝐶, define 𝑔(𝑞) = 𝜏 if and only if 𝛾(𝜏 ) = 𝑞. Then, the function 𝜏 : 𝐼 → 𝐽
is differentiable at 0. Consequently, the function 𝑔 : 𝐶 → 𝐽 is differentiable at
𝑝 and
𝑔(𝜎(ℎ)) − 𝑔(𝑝)
𝑑𝑔𝑝 (𝜎 ′ (0)) = lim = 𝜏 ′ (0).
ℎ→0 ℎ
Furthermore,
1
𝛾 ′ (0) = ′ 𝜎 ′ (0). (B.18)
𝜏 (0)
Proof: As in the proof of Lemma B.0.11, let 𝑗 ∈ {1, 2, . . . , 𝑛} be such that
where
𝐸1 (𝜏 (ℎ)) 𝐸2 (ℎ)
lim =0 and lim = 0. (B.21)
𝜏 (ℎ)→0 𝜏 (ℎ) ℎ→0 ℎ
We obtain from (B.20) and (B.19) that
[ ]
𝜏 (ℎ) 1 𝐸1 (𝜏 (ℎ)) 𝑥′𝑗 (0) 1 𝐸2 (ℎ)
1+ ′ = ′ + ′ ,
ℎ 𝑦𝑗 (0) 𝜏 (ℎ) 𝑦𝑗 (0) 𝑦𝑗 (0) ℎ
where 𝜎 : (−𝜀, 𝜀) → 𝐶 is any 𝐶 1 map defined on (−𝜀, 𝜀), for some 𝜀 > 0, such
that 𝜎 ′ (𝑡) ∕= 0 for all 𝑡 ∈ (−𝜀, 𝜀) and 𝜎(0) = 𝑝.
let 𝜀 > 0 be sufficiently small so that (𝑡−𝜀, 𝑡+𝜀) ⊂ 𝐽, and define 𝜎 : (−𝜀, 𝜀) → 𝐶
by
𝜎(𝜏 ) = 𝛾(𝑡 + 𝜏 ), for all 𝜏 ∈ (−𝜀, 𝜀).
Then, 𝜎 is a 𝐶 1 map satisfying 𝜎 ′ (𝜏 ) = 𝛾 ′ (𝜏 + 𝑡) ∕= 0 for all 𝜏 ∈ (−𝜀, 𝜀) and
𝜎(0) = 𝛾(𝑡). Observe also that 𝜎 ′ (0) = 𝛾 ′ (𝑡). It then follows by Definition
B.0.13 that 𝛾 ′ (𝑡) ∈ 𝑇𝛾(𝑡) 𝐶 for all 𝑡 ∈ 𝐽.
𝑑
[𝑔(𝛾(𝑡))] = 𝑑𝑔𝛾(𝑡) (𝛾 ′ (𝑡)), for all 𝑡 ∈ 𝐽. (B.23)
𝑑𝑡
Proof: Put 𝜎(ℎ) = 𝛾(𝑡 + ℎ), for ∣ℎ∣ sufficiently small. By Definition B.0.10,
where
∣𝐸𝛾(𝑡) (ℎ)∣
lim = 0. (B.25)
ℎ→0 ∣ℎ∣
The statement in (B.23) now follows from (B.24), (B.25), and the linearity of
the map 𝑑𝑔𝛾(𝑡) : 𝑇𝛾(𝑡) (𝐶) → ℝ.
Proof of Theorem B.0.7: Let 𝐼 and 𝐽 denote open intervals of real numbers con-
taining closed and bounded intervals [𝑎, 𝑏] and [𝑐, 𝑑], respectively, and 𝛾1 : 𝐼 →
ℝ𝑛 and 𝛾2 : 𝐽 → ℝ𝑛 be 𝐶 1 paths. Suppose that 𝐶 = 𝛾1 ([𝑎, 𝑏]) = 𝛾2 ([𝑐, 𝑑]) is a
𝐶 1 simple curve parametrized by 𝛾1 and 𝛾2 . Define 𝜏 : 𝐽 → 𝐼 by 𝜏 = 𝑔 ∘ 𝛾2 ,
where 𝑔 = 𝛾1−1 , the inverse of 𝛾1 . By Lemma B.0.12, 𝑔 : 𝐶 → 𝐼 is differentiable
on 𝐶. It therefore follows by the Chain Rule (Proposition B.0.14) that 𝜏 is
differentiable and
Taking norms on both sides of (B.26), and using the fact that 𝛾1 and 𝛾2 are
parametrizations, we obtain from (B.26) that
∥𝛾2′ (𝑡)∥
∣𝜏 ′ (𝑡)∣ = , for all 𝑡 ∈ 𝐽. (B.27)
∥𝛾1′ (𝜏 (𝑡))∥
113
Thus, either
𝜏 ′ (𝑡) > 0, for all 𝑡 ∈ 𝐽, (B.28)
or
𝜏 ′ (𝑡) < 0, for all 𝑡 ∈ 𝐽. (B.29)
If (B.28) holds true, then the proof of Theorem B.0.7 is complete. If (B.29) is
true, consider the function 𝜏˜ : 𝐽 → 𝐼 given by