2 Notes
2 Notes
4 k-forms 66
4.1 Differential forms revisited: an algebraic approach . . . . . . . . . . . . 66
4.2 Multiplying k-forms: the wedge product . . . . . . . . . . . . . . . . 73
4.3 Differentiating k-forms: the exterior derivative . . . . . . . . . . . . . 79
4.4 The exterior derivative and vector calculus . . . . . . . . . . . . . . . 90
4.5 Physical interpretation of grad, curl, div . . . . . . . . . . . . . . . . 99
4.6 Exact and closed k-forms . . . . . . . . . . . . . . . . . . . . . . 109
4.7 The pullback of a k-form . . . . . . . . . . . . . . . . . . . . . . 119
4.8 Hodge star . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
vii
CONTENTS viii
Appendices
there is always a “dx” inside the integrand? What does it mean, and why is it there? When
you were introduced to definite integrals in Calculus I, they were defined as limits of Riemann
sums: ⁄ n
b ÿ
f (x) dx = lim f (a + i x) x,
a næŒ
i=1
where x is the width of the rectangles in the Riemann sum. It was then argued that in the
limit as n æ Œ, the width of the rectangles goes to zero, and somehow x becomes “dx”,
which is some sort of infinitesimal width of the rectangles. But what does that mean, really?
What is this “dx” thing inside the definite integral?
And it’s not like we can just forget about it, even though many students actually do in
first year. :-) We all know how important dx is: just think of substitution for definite integrals.
Without the dx, substitution would fail miserably. It must be there. Why?
Things become even more interesting with double and triple integrals:
⁄⁄ ⁄⁄⁄
f dA, f dV,
D D
with dA = dxdy and dV = dxdydz in Cartesian coordinates. What are these objects dA and
dV , and why must they be there in the integrand?
What we will do in this course is provide an answer to this question. Our goal is to
define a unified theory of integration for curves, surfaces, and volumes, which will make
the appearance of dx, dA, and dV natural. To achieve this, we must define a new type of
objects that will play the role of integrands: those are called differential forms, or n-forms.
More specifically, one-forms are objects that can be integrated over curves, two-forms over
1
CHAPTER 1. A PREVIEW OF VECTOR CALCULUS 2
surfaces, and three-forms over volumes. In other words, it was all a big lie: what you should
be integrating is not functions, but rather differential forms!
But before we start, we can already identify two key guiding principles for the construction,
using what we already know about integration.
1.1.2 Vector calculus and differential forms: two sides of the same coin
What we will study in this course is known as vector calculus. There are two main approaches
to vector calculus. On the one hand, there is the “traditional” approach, which involves
concepts such as gradient, curl, div, etc. It relies heavily on the geometry of R3 , and is very
explicit. But at first it seems like a complicated amalgation of strange constructions and
definitions that satisfy all kinds of intricate identities. My recollections of learning vector
calculus is that it involves many formulae that appear to come out of nowhere and that one
needs to learn by heart. Not fun.
CHAPTER 1. A PREVIEW OF VECTOR CALCULUS 3
On the other hand, there is the “modern” approach, pioneered by Cartan, which relies
on the definition of differential forms. This approach is a little more abstract, but is much
more unified and elegant. It brings together all the concepts of vector calculus in a unified
formalism, from which all the identities and formulae come out naturally. It also does not
rely on the geometry of R3 , and is naturally generalized to Rn (even though we will focus on
R3 in this course). I remember this feeling of “ah, now this all makes sense!” when I learned
differential forms later on in my studies.
In this course we will take the perhaps non-traditional approach of introducing vector
calculus directly through the unified formalism of differential forms, guided by the exposition
of the previous subsection. The challenge is to make the concepts accessible to second-year
students, stripping them down from the fancier definitions of differential geometry. But I
truly believe that this is possible, and that it will make the whole theory of vector calculus
much more interesting and unified, and less reliant on brute force memorization.
But, at the same time, it is important for students to be fluent with the traditional
concepts of vector calculus. Indeed, students who will study topics like fluid mechanics,
electromagnetism, applied mathematics, etc. will repeadtedly encounter vector calculus,
usually expressed in the traditional language. Moreover, traditional concepts such as grad, div,
curl, are often useful for explicit calculations. So in this course we will translate all concepts
from differential forms to standard vector calculus every step of the way.
In the end, the goal is for students to be fluent with both approaches: to see the beauty
and elegance of differential forms, and to be able to use the traditional approach for explicit
calculations.
Chapter 2
In this section we study one-forms, which will become the objects that can be integrated along
curves. The counterpart to one-forms in traditional vector calculus is vector fields.
Objectives
You should be able to:
2.1.1 One-forms
Let us start by defining one-forms in R, R2 , R3 , and more generally Rn .
Definition 2.1.1 One-forms. A differential one-form (or simply one-form) on an open
interval (or union of open intervals) U ™ R is an expression of the form
Ê = f (x) dx
with f : U æ R a function with continuous derivatives (we say that the function f is “smooth”,
or C Œ , on U ).
A one-form on an open subset1 U ™ R2 is an expression of the form
Ê = f (x, y) dx + g(x, y) dy
4
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 5
with f, g : U æ R functions with continuous partial derivatives (again, we say that they are
smooth, or C Œ , on U ).
A one-form on an open subset U ™ R3 is an expression of the form
Of course a similar correspondence holds in Rn . This is the starting point for the dictionary
between the modern approach (one-forms) and the traditional approach (vector fields) to
vector calculus.
Example 2.1.4 A one-form and its associated vector field. The expression
Ê = x dx + x sin(y) dy
2.1.3 Exercises
1. Consider the vector field in R3 given by F(x, y, z) = (x, sin y, xyz). Write the correspond-
ing one-form.
Solution. The corresponding one-form is Ê = x dx + sin y dy + xyz dz.
2. Consider the vector field v(x, y) = (≠y, x) in R2 .
(a) Plot the given vector field at points with integer coordinates.
(b) If the vector field represents the velocity of a flowing fluid, what kind of motion is
this vector field describing?
Solution.
(a) We can plot the vector field by plotting arrows for points in R2 with integer
coordinates. For instance, at the point (1, 1), the vector field would be the vector
v(1, 1) = (≠1, 1). At (0, 1), we would get v(0, 1) = (≠1, 0). And so on and so forth.
The result is the following plot, which was obtained via this geogebra app.
(b) Looking at the plot, we see that the fluid is rotating around the origin counter-
clockwise. This type of fluid motion is called “vortex”.
3. Consider a vector field v(x, y) = (5, 1). Suppose that it is the velocity field of a moving
fluid. Suppose that you drop an object at the origin at time t = 0. Where will the object
be at time t = 2?
Solution. The velocity field v(x, y) = (5, 1) is constant, which means that all points in
the fluid are moving at the same constant velocity. In other words, each point in the
fluid is going through a linear motion of the form
where x0 is some initial position. If we drop an object at the origin at t = 0, its position
vector is then described by x(t) = t(5, 1) since its initial position is x0 = (0, 0). We
conclude that at time t = 2, the object is at position x(2) = 2(5, 1) = (10, 2).
4. Consider the object Ê = ≠ x2 +y
y
2 dx +
x
x2 +y 2
dy. Is this a one-form on R2 ?
Solution. Ê is not a one-form on R2 , because its component functions are not defined
at the origin (0, 0) œ R2 (we would be dividing by zero). However, it is a one-form on
the open subset U = R2 \ {(0, 0)} where we removed the origin,2 since the component
functions are smooth on U .
Objectives
You should be able to:
• Rephrase the statement as the "screening test" for conservative vector fields.
• Use the screening test to show that a given vector field cannot be conservative.
2
The notation R2 \ {(0, 0)} means “R2 minus the point (0, 0)”.
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 8
ˆf ˆf ˆf
df = dx + dy + dz.
ˆx ˆy ˆz
ˆf ˆf
df = dx + dy = 2xey dx + x2 ey dy,
ˆx ˆy
GM m
Ê=≠ (x dx + y dy + z dz) .
r3
It is easy to see that the gravitational force field is conservative, or equivalently that the
one-form Ê is exact, since Ê = dÏ with the function Ï given by
GM m
Ï(x, y, z) = .
r
Ï is the potential function, which from physics you may recognize as minus the gravitational
potential energy. ⇤
back to it in Section 3.6. For the time being, we will be able to find a necessary condition for a
one-form to be exact, which in the context of vector calculus is sometimes called a "screening
test" for conservative vector fields.
Let us focus first on one-forms and vector fields on U ™ R2 .
Definition 2.2.9 Closed one-forms in R2 . We say that a one-form Ê = f dx + g dy on
U ™ R2 is closed if
ˆf ˆg
= .
ˆy ˆx
⌃
This definition may seem ad hoc, but it is important because of the following lemma.
Lemma 2.2.10 Exact one-forms in R2 are closed. If a one-form Ê on U ™ R2 is exact,
then it is closed.
Proof. Suppose that Ê is exact: then it can be written as
ˆF ˆF
Ê = f dx + g dy = dF = dx + dy
ˆx ˆy
ˆf ˆ2F
=
ˆy ˆyˆx
is equal to
ˆg ˆ2F
= .
ˆx ˆxˆy
Equality of the two expressions follows from the Clairaut-Schwarz theorem, which states that
partial derivatives commute, as long as they are continuous. But continuity of the partial
derivatives is guaranteed by the fact that all partial derivatives of F exist and are continuous,
since by definition (see Definition 2.1.1) one-forms are assumed to have smooth component
functions. Therefore the one-form is closed. ⌅
Note however that the converse statement is not necessarily true: not all closed one-forms
on U ™ R2 are exact. In fact, the question of when closed one-forms are exact is an important
one; the result is known as Poincare’s lemma. We will come back to this in Section 3.6. But
what Lemma 2.2.10 tells us is that one-forms that are not closed cannot be exact.
There is of course an analogous statement for conservative vector fields. The only minor
difference is that we need to impose a condition on the component functions of vector fields,
since vector fields are not always smooth by definition (see Definition Definition 2.1.2.
Lemma 2.2.11 Screening test for conservative vector fields in R2 . If a vector field
F = (f1 , f2 ) on U ™ R2 is conservative and has component functions that are continuously
differentiable, then it passes the screening test:
ˆf1 ˆf2
= .
ˆy ˆx
Proof. Same as for Lemma 2.2.10, but for the associated vector fields. ⌅
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 11
In the context of vector fields, it is called a screening test, because it is a quick test to
determine whether a vector field has a chance at all to be conservative. In other words, if a
vector field does not pass the screening test, then it is certainly not conservative. However,
if it passes the screening test, at this stage we cannot conclude anything. Just like closed
one-forms are not necessarily exact.
Example 2.2.12 Exact one-forms are closed. Consider the exact one-form on R2 from
Example 2.2.7, Ê = cos x dx ≠ sin y dy. We show that it is closed, according to Definition 2.2.9.
The partial derivatives are easily calculated:
ˆ ˆ
cos x = 0, (≠ sin y) = 0.
ˆy ˆx
Thus Definition 2.2.9 is satisfied, and Ê is closed. ⇤
Example 2.2.13 Closed one-forms are not necessarily exact. If a vector field is
conservative, then it passes the screening test. Correspondingly, if a one-form is exact, then
it is closed. But the converse statement is not necessarily true (we will revisit it later in
Section 3.6). Consider for instance the one-form
y x
Ê=≠ dx + 2 dy.
x2 + y 2 x + y2
Calculating the partial derivatives for Definition 2.2.9, we get that
3 4 3 4
ˆ y y 2 ≠ x2 ˆ x y 2 ≠ x2
≠ 2 = , = .
ˆy x + y2 x2 + y 2 ˆx x + y2
2 x2 + y 2
The two expressions are equal, and thus Ê is closed. However, one can show that Ê is not
exact: there does not exist a function f such that Ê = df (see Exercise 3.4.3.2). ⇤
2.2.5 Exercises
1. Consider the function f (x, y, z) = sin(y) + zx. Find its differential df .
Solution. df is given by:
ˆf ˆf ˆf
df = dx + dy + dz
ˆx ˆy ˆz
=z dx + cos(y) dy + x dz.
2. Consider the one-form Ê = x dx ≠ y dy on R2 . Show that it is exact, and find a function
f such that Ê = df .
Solution. Suppose that there exists a function f (x, y) such that df = ˆf
ˆx dx + ˆf
ˆy dy =
x dx ≠ y dy. Then we must have
ˆf ˆf
= x, = ≠y.
ˆx ˆy
Integrating the first equation (recalling that this is a partial derivative, so the “constant
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 13
x2
f (x, y) = + g(y)
2
for some function g(y). Substituting into the second equation, we get
ˆf
= g Õ (y) = ≠y, ,
ˆy
which can be integrated to
y2
g(y) = ≠ +C
2
for any constant C. Since we are only interested in finding one f such that df = Ê, we
can choose C = 0. We conclude that the function
x2 y 2
f (x, y) = ≠
2 2
is such that df = Ê, and hence that Ê is an exact one-form.
3. True or False. If F and G are both conservative, then F + G is also conservative.
Solution. True. If F and G are both conservative, then F = Òf and G = Òg for some
functions f and g. But then F + G = Ò(f + g), and hence F + G is also conservative.
4. Consider the vector field F(x, y, z) = (yzexy , xzexy , exy ). Show that it is conservative
and find a potential.
Solution. F is conservative if there exists a function f such that F = Òf . So we need
to solve the equations
ˆf ˆf ˆf
= yzexy , = xzexy , = exy .
ˆx ˆy ˆz
Let us start by integrating the last one. We get:
for some function g(x, y). Substituting back into the second one, we get:
ˆf ˆg
= xzexy + = xzexy ,
ˆy ˆy
f (x, y, z) = zexy .
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 14
5. Show that the one-form Ê = f Õ (x) dx + g Õ (y) dy on R2 is exact, for any smooth functions
f, g : R æ R.
Solution. Consider the function F (x, y) = f (x) + g(y). Its differential is
ˆF ˆF
dF = dx + dy = f Õ (x) dx + g Õ (y) dy.
ˆx ˆy
ˆF ˆF
= cos(y), = ≠x sin(y).
ˆx ˆy
Integrating the first equation, we get F (x, y) = x cos(y) + h(y) for some function h(y).
Substituting in the second equation, we get
ˆF
= ≠x sin(y) + hÕ (y) = ≠x sin(y),
ˆy
while
ˆg
= 1.
ˆx
Thus Ê is not closed, and hence it cannot exact.
9. Consider the vector field F(x, y) = (xy n , 2x2 y 3 ) for some positive integer n. Find the
value of n for which F is conservative on R2 , and find a potential for this value of n.
Solution. For F to be conservative, it must pass the screening test. We write F =
(f1 , f2 ), and calculate:
ˆf1 ˆf2
= nxy n≠1 , = 4xy 3 .
ˆy ˆx
We see that the two partial derivatives are equal for all (x, y) if and only if n = 4, in which
case the vector field becomes F = (xy 4 , 2x2 y 3 ). To show that it is conservative and find
a potential function, we are looking for a function f (x, y) such that Òf = (xy 4 , 2x2 y 3 ).
So we must have
ˆf ˆf
= xy 4 , = 2x2 y 3 .
ˆx ˆy
Integrating the first equation, we get
1
f (x, y) = x2 y 4 + g(y)
2
for some function g(y). Substituting in the second equation, we get
ˆf
= 2x2 y 3 + g Õ (y) = 2x2 y 3 ,
ˆy
Objectives
You should be able to:
dG d dF dx
= (F (x(t)) = .
dt dt dx dt
So by changing variable from x to t, we get a new one-form, let’s call it ÷, defined by
3 4
dG dF dx
÷= dt = dt.
dt dx dt
We now generalize this to all one-forms on R (or open subsets thereof), not just exact
one-forms. Given a one-form Ê = f (x) dx written in terms of a variable x, if we think of
x = x(t) as a function of a new variable t, then by changing variable from x to t we get a new
one-form ÷: 3 4
dx
÷ = f (x(t)) dt.
dt
This defines how one-forms transform under changes of variables. As we will see, this
transformation property is what lies behind the substitution formula for definite integrals.
Remark 2.3.1 The upshot of this brief discussion is that it is easy to remember how one-forms
in R transform under changes of variables. If we write Ê = f (x) dx, and do a change of
variable x = x(t), then all we need to do is rewrite the coefficient function as a function of t
by composition f (x(t)), and then “transform the differential” dx as dx(t) = dxdt dt. This gives
us the new one-form ÷ = f (x(t)) dt dt.
dx
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 17
g = f ¶ „ : V æ R.
„ú f := f ¶ „ : V æ R.
d„
„ú Ê = f („(t)) dt = (sin t)2 cos t dt.
dt
This is of course the same thing as implementing the change of variables x æ t in the one-form
Ê. ⇤
2.3.3 Exercises
1. Consider the one-form Ê = ex sin(x) dx on R, and the smooth function „ : R>0 æ R
with „(t) = ln(t) (where R>0 is the set of positive real numbers). Find the pullback
one-form „ú Ê. Where is the one-form „ú Ê defined?
Solution. First, since „ : R>0 æ R, and Ê is a one-form defined on all of R, by
definition of the pullback we see that the pullback one-form „ú Ê is defined only on R>0 ,
i.e. for all positive real numbers. Using the definition of pullback, we find its expression
as:
d„
„ú Ê =e„(t) sin(„(t)) dt
dt
1
=eln(t) sin(ln(t)) dt
t
1
=t sin(ln(t)) dt
t
= sin(ln(t)) dt.
2. Consider the one-form Ê = x1 dx defined on R>0 , and the smooth function „ : R æ R>0
defined by „(t) = et . Find the pullback one-form „ú Ê.
Solution. First, by definition of the pullback we see that „ú Ê is defined on all of R.
We find its expression to be:
1 Õ
„ú Ê = „ (t) dt
„(t)
=e≠t et dt
=dt.
„ú (df ) = d(„ú f ).
In other words, the pullback commutes with the exterior derivative of a function.
Solution. From the definition of the pullback of a function, we can write the right-
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 19
hand-side as:
df („(t))
d(„ú f ) = d(f („(t)) = dt.
dt
Using the chain rule, this can be written as
d„
d(„ú f ) =f Õ („(t)) dt
dt
=„ú (f Õ (x) dx)
=„ú (df ),
where we used the definition of the pullback of a one-form. This concludes the proof.
Objectives
You should be able to:
• State and use the three fundamental properties of the pullback of one-forms.
Let Ê be a one-form on U ™ Rn , where n is a positive integer. We now consider a smooth
function „ : V æ U , where V ™ Rm , with again m a positive integer. Note that m and n
don’t have to be the same: we could have, say U ™ R3 , and V ™ R2 . Our goal is to define the
pullback „ú Ê, which should be a one-form on V .
Just to be concrete: we could take, for instance, a one-form Ê on R3 , and a smooth
function „ : R2 æ R3 . The pullback „ú Ê should then be a one-form on R2 .
Note that our notion of pullback should generalize the definition of pullback in Defini-
tion 2.3.4, which should consist in the case with m = n = 1.
„ú f (t) = f (t, t2 , 1) = t3 + 1.
⇤
Example 2.4.3 The pullback of a function from R3to R2 .
We can do the same thing
but pulling back to R instead of R. Suppose that f is a smooth function on R3 , that is
2
3. „ú (df ) = d(„ú f ).
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 21
It turns out that this is completely sufficient to fully determine the pullback of any one-form.
Let us see why.
Lemma 2.4.4 The pullback of dx. Let dx be the basic one-form on R3 . Let „ : V æ R3
be a smooth function, where V ™ Rm is an open subset, with m a positive integer. Write
t = (t1 , . . . , tm ) for an m-dimensional vector in V , and „(t) = (x(t), y(t), z(t)). Then
m
ÿ ˆx
„ú (dx) = dti .
i=1
ˆti
From Definition 2.4.1, we can calculate „ú f . We get „ú f (t) = x(t). We thus obtain
m
ÿ ˆx
„ú (dx) = dx(t) = dti ,
i=1
ˆti
„ú Ê = „ú (f dx + gdy + hdz) = („ú f )„ú (dx) + („ú g)„ú (dy) + („ú h)„ú (dz).
We then use Lemma 2.4.4 to evaluate „ú (dx), „ú (dy) and „ú (dz), and from Definition 2.4.1
we know that „ú f (t) = f („(t)), and similarly for g and h. ⌅
Example 2.4.6 The pullback of a one-form from R3 to R. Suppose that Ê =
f dx + gdy + hdz is a one-form on R3 . Let „ : R æ R3 be a smooth function, which takes a
point in R and maps it to a vector in R3 . We write „(t) = (x(t), y(t), z(t)). Then the pullback
„ú Ê is a one-form on R given by:
3 4
dx dy dz
„ú Ê = f („(t)) + g(„(t) + h(„(t)) dt.
dt dt dt
For instance, if Ê = xdx + xydy + z 2 dz, and „(t) = (x(t), y(t), z(t)) = (t2 , t, 1), then
3 4
dx dy dz
„ú Ê = x(t) + x(t)y(t) + z(t)2 dt
dt dt dt
1 2
= (t2 )(2t) + (t2 )(t)(1) + (1)(0) dt
=3t3 dt.
⇤
Example 2.4.7 The pullback of a one-form from to R3 R2 .
Suppose that Ê =
f dx + gdy + hdz is a one-form on R3 . Let „ : R2 æ R3 be a smooth function, which takes a
point in R2 and maps it to a vector in R3 . We write „(t) = (x(t), y(t), z(t)), with t = (t1 , t2 ).
Then the pullback „ú Ê is a one-form on R2 given by:
3 4 3 4
ˆx ˆx ˆy ˆy
„ú Ê =f („(t)) dt1 + dt2 + g(„(t)) dt1 + dt2
ˆt1 ˆt2 ˆt1 ˆt2
3 4
ˆz ˆz
+ h(„(t)) dt1 + dt2 .
ˆt1 ˆt2
For instance, if Ê = xdx + xydy + z 2 dz, and „(t) = (x(t), y(t), z(t)) = (t1 t2 , t2 , t1 + t2 ),
then
3 4 3 4
ˆx ˆx ˆy ˆy
„ú Ê =x(„(t)) dt1 + dt2 + x(„(t))y(„(t)) dt1 + dt2
ˆt1 ˆt2 ˆt1 ˆt2
3 4
ˆz ˆz
+ z(„(t))2 dt1 + dt2 .
ˆt1 ˆt2
=(t1 t2 )(t2 dt1 + t1 dt2 ) + (t1 t2 )(t2 )(dt2 ) + (t1 + t2 )2 (dt1 + dt2 )
=(t1 t22 + (t1 + t2 )2 )dt1 + (t21 t2 + t1 t22 + (t1 + t2 )2 )dt2 .
⇤
Example 2.4.8 Consistency check: the pullback of a one-form from R to R.
As a consistency check, we show that the pullback of a one-from from R to R reduces to
Definition 2.3.4. Let Ê = f (x)dx on U ™ R, and „ : V æ U with V ™ R. We write „(t) = x(t).
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 23
which indeeds reproduces Definition 2.3.4 with our notation „(t) = x(t). ⇤
We now have a very general definition of the pullback of a one-form. This will turn out to
be very useful to define the integral of a one-form, which is what we now turn to.
2.4.3 Exercises
1. Consider the function f : R2 æ R given by f (x, y) = ex+y + x + y, and the function
„ : R3 æ R2 given by „(u, v, w) = (u + v, v + w). Find the pullback „ú f . What is its
domain?
Solution. First, as f : R2 æ R and „ : R3 æ R2 , we see that the composition
„ú f = f ¶ „ : R3 æ R2 æ R, i.e. the pullback „ú f is a function from R3 to R. So its
domain is R3 .
We calculate its expression by composition:
„ú f (u, v, w) =f (u + v, v + w)
=e(u+v)+(v+w) + (u + v) + (v + w)
=eu+2v+w + u + 2v + w.
2. Consider the one-form Ê = x2 dx on R, and the function „ : R2 æ R given by „(u, v) = u,
which projects on the u-axis. Find the pullback one-form „ú Ê on R2 . Interpret the result.
Solution. Let us write Ê = f dx = x2 dx. By the definition of pullback, we get:
3 4
ˆ„ ˆ„
„ Ê =f („(u, v))
ú
du + dv
ˆu ˆv
=u2 (1 du + 0 dv)
=u2 du.
We see that the pullback one-form looks the same, but written in terms of u instead
of x. However, „ú Ê is defined on R2 , while Ê was defined on R. Since the function „
here simply projects on the u-axis, what the pullback does here is extend the one-form
uniformly in the v-coordinate on the uv-plane; at any two points (u, v1 ) and (u, v2 ), the
one-form will be the same. Conceptually, this is what happens when we pullback using
a “forgetful map”, i.e. a map that somehow “forgets” some information (in this case,
the v-coordinate). The pullback then extends the object uniformly across the forgotten
structure.
3. Consider the one-form Ê = x2 dx + y 2 dy on R2 , and the map : R2 æ R2 with
(r, ◊) = (r cos ◊, r sin ◊), which defines polar coordinates. Find the pullback ú Ê.
Solution. We write Ê = f dx + g dy = x2 dx + y 2 dy, and (r, ◊) = (x(r, ◊), y(r, ◊)).
Then:
3 4 3 4
ˆx ˆx ˆy ˆy
ú
Ê =f ( (r, ◊)) dr + d◊ + g( (r, ◊)) dr + d◊
ˆr ˆ◊ ˆr ˆ◊
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 24
The notion of pullback allows us to easily calculate how one-forms change under
changes of coordinates, such as going from Cartesian to polar coordinates in this case.
2 dz
4. Consider the one-form Ê = Ôz on U = {(x, y, z) œ R3 | (x, y, z) ”= (0, 0, z)} (this is R3
x2 +y 2
with the z-axis removed), and the function „ : V æ U with „(r, ◊, ’) = r(cos ◊, sin ◊, ’),
and V = {(r, ◊, ’) œ R3 | r ”= 0}. Determine the pullback one-form „ú Ê.
Solution. By definition of the pullback, we get:
3 4
1 ˆ ˆ ˆ
„ Ê = r2 ’ 2
ú
(r’) dr + (r’) d◊ + (r’)d’
r ˆr ˆ◊ ˆ’
=r’ 2 (’ dr + r d’).
5. Let Ê be a one-form on R3 , and Id : R3 æ R3 the identity function defined by Id(x, y, z) =
(x, y, z). Show that Idú Ê = Ê.
Solution. Write Ê = f dx + g dy + h dz. By definition of the pullback, we get:
3 4
ˆ ˆ ˆ
Idú Ê =f (x, y, z) (x) dx + (x) dy + (x) dz
ˆx ˆy ˆz
3 4
ˆ ˆ ˆ
+ g(x, y, z) (y) dx + (y) dy + (y) dz
ˆx ˆy ˆz
3 4
ˆ ˆ ˆ
+ h(x, y, z) (z) dx + (z) dy + (z) dz
ˆx ˆy ˆz
=f (x, y, z) dx + g(x, y, z) dy + h(x, y, z) dz
=Ê,
In other words, it doesn’t matter whether we pullback in one or two steps through the
„
chain of maps W æ V æ U .
–
We note here that while the exercise is only asking you to prove it for open subsets
U, V, W ™ R, this property is true in general, not just in R.
Solution. Let us write Ê = f (x) dx, „ = „(t), and – = –(u). On the one hand, we
have:
d
(„ ¶ –)ú Ê = f („(–(u))) („(–(u))) du.
du
On the other hand, we have
d
„ú Ê = f („(t)) („(t)) dt,
dt
CHAPTER 2. ONE-FORMS AND VECTOR FIELDS 25
and 3 4-
d - d
– („ Ê) = f („(–(u))
ú ú
„(t) - –(u) du.
dt t=–(u) du
But 3 4-
d d - d
(„(–(u))) = „(t) - –(u)
du dt t=–(u) du
We study how one-forms can be integrated along curves, which leads to the definition of
(oriented) line integrals (also called “work integrals”). An important result in this section
is the Fundamental Theorem of line integrals, which is the natural generalization of the
Fundamental Theorem of Calculus to integrals of one-forms along curves in R2 and R3 .
Objectives
You should be able to:
• Determine how the integral changes when reversing the orientation of the interval.
• Relate the transformation property of one-forms to the substitution formula for definte
integrals.
where on the right-hand-side we use the standard definition of definite integrals from calculus.
⌃
26
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 27
Well, that was simple! This is just the standard notion of definite integrals from calculus.
All we are doing is introducing some fancy notation for it. Great!
Example 3.1.2 An example of an integral of a one-form over an interval. Consider
the one-form Ê = x3 dx on R. Suppose that you want to integrate it over the interval [0, 1].
Then the integral is ⁄ ⁄ 1
1
Ê= x3 dx = ,
[0,1] 0 4
where we used the Fundamental Theorem of Calculus to evaluate the integral as usual, since
we are back in the realm of the definite integrals that we know and love. ⇤
What is interesting however is to study what our guiding principles of reparametrization-
invariance and orientability become in this simple context. Let us start with orientability.
⌃
When the orientation is canonical, that is [a, b]+ = [a, b], we recover our original definition
Definition 3.1.1. But when the interval is oriented in the direction of decreasing real numbers,
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 28
since this is nothing else but the substitution formula for definite integrals that you proved in
Calculus I! Indeed, since a = „(c) and b = „(d), we can rewrite this equation as
⁄ d ⁄ „(d)
d„
f („(t)) dt = f (x) dx,
c dt „(c)
Proof. From the substitution formula for definite integrals, we know that
⁄ d ⁄ „(d) ⁄ a ⁄ b
d„
f („(t)) dt = f (x) dx = f (x) dx = ≠ f (x) dx.
c dt „(c) b a
s
We recognize the right-hand-side as ≠ [a,b] Ê, which completes the proof. ⌅
It is thus not true that the integral is invariant under all reparametrizations: it is only
invariant under reparametrizations that perserve the orientation of the interval. Under
reparametrizations that reverse the orientation, the integral changes sign, as expected.
We have thus shown that the integral of a one-form over an interval is both oriented and
reparametrization-invariant, our two guiding principles. In the context of definite integrals,
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 30
orientability reduces to the statement that definite integrals pick a sign when we exchange
limits of integration, and reparametrization-invariance to the substitution formula. Neat!
3.1.5 Exercises
2
1. Consider the one-form Ê = xex dx. Compute the integral of Ê over the interval [1, 2],
with both choices of orientation.
s
Solution. We first calculate [1,2] Ê with the canonical choice of orientation in the
direction of increasing real numbers, i.e. from 1 to 2. We get:
⁄ ⁄ 2 ⁄ 4
x2 1 1
Ê= xe dx = eu du = (e4 ≠ e),
[1,2]+ 1 2 1 2
where we used the substitution u = x2 . As for the other choice of orientation in the
direction of decreasing real numbers, i.e. from 2 to 1, we get:
⁄ ⁄ 1 ⁄ 1
x2 1 1
Ê= xe dx = eu du = ≠ (e4 ≠ e),
[1,2]≠ 2 2 4 2
Solution. The map „(t) = can be restricted to the interval [0, 3] µ R. We see that
et
its image is the interval [1, e3 ], and that the map is injective. Moreover, „(0) = 1 and
„(3) = e3 , so it preserves orientation. Thus we know that
⁄ ⁄
„ú Ê = Ê.
[0,3] [1,e3 ]
Objectives
You should be able to:
– : [a, b] æ Rn
t ‘æ –(t) = (x1 (t), . . . , xn (t)),
such that:
3. if –(s) = –(t) for any two distinct s, t œ [a, b], then s, t œ {a, b}. 2 In other words, the
map – is injective everywhere except possibly at the end points a and b.
that a parametrization induces a well defined orientation on the curve, as we will see.
Property 3 in Definition 3.2.1 imposes that the map – is injective everywhere except
possibly at the endpoints a and b of the interval, which could be mapped to the same point.
This ensures that the image curve does not cross itself, and that the parametrization does not
go through the same path multiple times (for instance, going around a circle two times).
Because of Property 3, we can distinguish between two types of parametric curves,
depending on whether the image curve has endpoints or not.
Definition 3.2.2 Closed parametric curves. Let – : [a, b] æ Rn be a parametric curve,
with image C = –([a, b]) µ Rn .
1. If –(a) = –(b), then we say that the parametric curve is closed, as the image curve has
no endpoints (it is a loop).
2. If – : [a, b] æ Rn is injective, we call the set ˆC = {–(a), –(b)} consisting only in the
endpoints of C the boundary of the curve, which we denote by ˆC.
⌃
Example 3.2.3 Parametrizing the unit circle. Consider the function – : [0, 2fi] æ R2
given by
–(◊) = (cos ◊, sin ◊).
It is easy to see that the image –([a, b]) µ R2 is the unit circle x2 + y 2 = 1. Let us check that
the conditions in Definition 3.2.1 are satisfied:
2. –Õ (t) = (≠ sin ◊, cos ◊) is never zero over the interval [0, 2fi].
3. The only two values of ◊ œ [0, 2fi] that have the same image are ◊ = 0 and ◊ = 2fi, the
endpoints of the interval.
Because of the last statement above, this is an example of a closed parametric curve. ⇤
In other words, we simply differentiate the component functions of the vector-valued function
–. As T is a vector-valued function, it assigns a vector to every point on the curve C, namely
the tangent vector to the curve at that point. ⌃
We know that its image is the unit circle x2 + y 2 = 1. What is the induced orientation on the
circle? We calculate the tangent vector:
Now consider ◊ = 0. The parametrization maps this point to –(0) = (1, 0), so this is the point
with coordinates (1, 0) on the unit circle. As for the tangent vector, we see that T(0) = (0, 1),
and hence the tangent vector at the point (1, 0) on the unit circle is pointing upwards. This
tells us that we are moving along the curve in a counterclockwise direction, which is the
induced orientation on the unit circle. ⇤
„ú – : [c, d] æ Rn
u ‘æ(„ú x1 (u), . . . , „ú xn (u)) = (x1 („(u)), . . . , xn („(u)))
and by assumption d„du never vanishes on [c, d]. As for property three, it follows since „ is
assumed to be injective. ⌅
Remark 3.2.10 We note that since d„ du is continuous, and is never zero on [c, d], then it is
either everywhere positive on [c, d], or everywhere negative.
Since a parametric curve induces a choice of orientation on the curve, and there are only two
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 36
orientation-preserving. If d„ du < 0 for all u œ [c, d], then the two parametrizations – and
„ – induce opposite orientations, and we call the reparametrization orientation-reversing.
ú
Proof. We simply need to compare the tangent vectors. Let T– be the tangent vector associated
to the parametrization –, and T„ú – be the tangent vector associated to „ú –. We have:
d„ ! Õ " d„
T„ú – (u) = x1 („(u)), . . . , xÕn („(u)) = T– („(u)).
du du
fi fi
„ú –(t) = (cos( ≠ t), sin( ≠ t)) = (sin t, cos t),
2 2
which is our second parametrization —. Since „Õ (t) < 0, we expect – and — = „ú – to induce
opposite orientation, which is exactly what we observed. ⇤
If a piecewise parametric curve is the union of a number of parametric curves, and each
parametric curve is smooth, we call the piecewise parametric curve piecewise smooth.
Remark 3.2.13 We add one more piece of notation. It will be useful to distinguish between
curves that have self-intersection and those that do not. We say that a curve that doesn’t
intersect itself (except possibly at the endpoints) is simple. With our definition of parametric
curves Definition 3.2.1, the image of a parametric curve will be always be simple, as it cannot
self-intersect.
Non-simple curves can be studied using piecewise parametric curves, as any non-simple
curve can be broken into a number of simple curves.
Example 3.2.14 Parametrizing a triangle. Consider the triangle with vertices A = (0, 0),
B = (0, 1) and C = (1, 0). We cannot parametrize it as a single parametric curve according to
Definition 3.2.1, since we cannot find a parametrization – that has a non-vanishing tangent
vector at the vertices of the triangle. Instead, we split it into the union of three parametric
curves, corresponding to the three edges of the triangle:
–1 parametrizes the edge AB, –2 the edge BC, and –3 the edge CA. As each parametric
curve is smooth, the union defines a piecewise smooth parametric curve. ⇤
3.2.6 Exercises
1. Find a parametrization for the straight line between the points (0, 1, 1) and (2, 3, 3) in
R3 .
Solution. We are given two points on the line. The vector d whose direction is parallel
to the line is given by d = (2, 3, 3) ≠ (0, 1, 1) = (2, 2, 2). So we can write an equation for
the line as
r(t) = (0, 1, 1) + t(2, 2, 2), 0 Æ t Æ 1.
In the language of this section, this gives us a parametrization of the line, in the form of
a map:
– :[0, 1] æ R3
t ‘æ –(t) = (2t, 1 + 2t, 1 + 2t).
2. Express the upper half of a circle of radius 3 and centered at the point (1, 0) as a
parametric curve. Determine the orientation induced by your parametrization.
Solution. The circle of radius 3 and centered at the point (1, 0) has equation
(x ≠ 1)2 + y 2 = 9.
It is easy to find a parametrization for the circle. We can take, for instance,
We want only the upper half of the circle though, so we need to restrict to y Ø 0. This
amounts to restricting the domain of our parametrization to 0 Æ t Æ fi. The resulting
parametrization can be written as the map – : [0, fi] æ R2 with –(t) = (3 cos t + 1, 3 sin t).
What is the induced orientation? The tangent vector to our parametric curve is
T(t) = (≠3 sin t, 3 cos t). Pick a point on the circle, say (4, 0). This corresponds to t = 0.
At this point, the tangent vector is T(0) = (0, 3), which is pointing upwards. This means
that our parametrization induces a counterclockwise orientation around the circle.
3. Consider the parametric curve – : [0, 4fi] æ R3 , –(t) = (cos t, sin t, 4t). What is the shape
of the image curve C = –([0, 4fi]) µ R3 ? What is the induced orientation?
Solution. Let us write –(t) = (x(t), y(t), z(t)). We see that x2 (t) + y 2 (t) = cos2 t +
sin2 t = 1, so all points on the curve C lie on the cylinder x2 + y 2 = 1 with radius 1.
Moreover, as t increases, z(t) = 4t increases linearly. So the curve is an helix on the
cylinder with radius 1 centred around the z-axis.
The curve starts at the point –(0) = (1, 0, 0) on the cylinder, and ends at the point
–(4fi) = (1, 0, 16fi). As t runs from 0 to 4fi, we see that the curve goes twice around
the cylinder. Moreover, the tangent vector is T(t) = (≠ sin t, cos t, 4); in particular, its
z-coordinate is always positive, which means that the tangent vector is always pointing
upwards. We conclude that the induced orientation is going upwards along the helix.
4. Suppose that a particle is moving along the parametric curve – : [0, fi] æ R3 with
–(t) = (sin(t2 ), cos(t), t). Find its velocity at t = fi.
Solution. The velocity vector is v(t) = (2t cos(t2 ), ≠ sin(t), 1). At t = fi, we get
v(fi) = (2fi cos(fi 2 ), 0, 1). This is the velocity of the particle at t = fi, which is of course
a vector as the particle is moving in three-dimensional space.
5. Consider the curve that is the intersection of the cylinder x2 + y 2 = 1 and the surface
z = x2 ≠ y 2 in R3 . Find a parametrization for the curve. Is it a closed curve?
Solution. A point on the cylinder x2 + y 2 = 1 can be parametrized by (x, y, z) =
(cos t, sin t, z), with 0 Æ t < 2fi and z œ R. But we want the curve to lie on the
surface z = x2 ≠ y 2 , so z is fixed as z = cos2 t ≠ sin2 t. We thus get a parametric
expression for a point on the curve as (x(t), y(t), z(t)) = (cos t, sin t, cos2 t ≠ sin2 t), with
0 Æ t < 2fi. To rewrite this as a parametric curve in the language of this section,
we include the endpoint t = 2fi. We get the parametric curve – : [0, 2fi] æ R3 with
–(t) = (cos t, sin t, cos2 t ≠ sin2 t). This is a closed curve, since –(0) = (1, 0, 1) and
–(2fi) = (1, 0, 1), i.e. the starting and ending points coincide.
6. Consider the map – : [≠1, 1] æ R2 with –(t) = (x(t), y(t)), where
Y
]t2 for 0 Æ t Æ 1
x(t) = , y(t) = t2 .
[≠t2 for ≠ 1 Æ t < 0
Show that it is not a parametric curve, according to Definition 3.2.1. What does the
image C = –([≠1, 1]) µ R2 look like?
Solution. At first one may think that this is valid parametric curve. The map – is
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 39
injective, so Property 3 is satisfied. Since its derivative is –Õ (t) = (xÕ (t), y Õ (t)) with
Y
]2t for 0 Æ t Æ 1
xÕ (t) = , y Õ (t) = 2t,
[≠2t for ≠ 1 Æ t < 0
–Õ (t) exists and is continuous for all t œ R, and thus Property 1 is also satisfied (note
that – cannot be extended to a smooth function however, since xÕ (t) is not differentiable
at t = 0, but that’s ok, it doesn’t have to). However, the problem is with Property 2: we
see that –Õ (0) = (0, 0), with 0 œ [≠1, 1]; thus Property 2 is not satisfied. We conclude
that this is not a parametric curve.
The image curve C = –([≠1, 1]) is the set of points in R2 satisfying the equation
y = |x| between (≠1, 1) and (1, 1), which has a corner at the origin. This example
highlights one of the reasons why Property 2 is there in Definition 3.2.1. If it wasn’t
there, this means that we could find a parametrization for the curve y = |x|; but we do
not want the image curve to have kinks or corners. Property 2 ensures that we cannot
find a parametrization for the whole curve y = |x| between (≠1, 1) and (1, 1) at once.
This is not to say however that we cannot deal with this curve. Just like for the
triangle, the idea is to consider it as a piecewise parametric curve. I.e., we realize the
line segment from (≠1, 1) to (0, 0) as a (smooth) parametric curve, the line segment from
(0, 0) to (1, 1) as another (smooth) parametric curve, and we take the union of the two
parametric curves.
Objectives
You should be able to:
• Define the line integral of a one-form along a parametric curve in R2 and R3 , and
evaluate it.
• Rewrite the definition of line integrals in terms of the associated vector fields.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 40
• Show that line integrals change sign under reparameterizations of the curve that reverse
its orientation.
⇤
As is becoming customary, we can translate between the language of differential forms and
the language of vector fields. If F is the vector field associated to Ê, and – is a parametric
curve in Rn , then the pullback one-form –ú Ê can be written as
–ú Ê = (F(–(t)) · T(t)) dt,
where T(t) is the tangent vector to the parametric curve, and · denotes the dot product of
vectors.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 41
5
=≠ + 2 sin(1) + 2 cos(1).
3
⇤
Remark 3.3.4 Line integrals over piecewise parametric curves. We note that we
can easily generalize the definition of line integrals to piecewise parametric curves, as in
Subsection 3.2.5. If the parametric curve is defined as a union of parametric curves, then to
integrate along the curve we simply add up the integrals over the pieces.
As for the second integral, we are integrating over the parametric curve – ¶ „ : [c, d] æ [a, b] æ
Rn . So we can write ⁄ ⁄
Ê= (– ¶ „)ú Ê.
„ú – [c,d]
„
However, from Exercise 2.4.3.6 we know that pulling back through the chain of maps [c, d] æ
[a, b] æ Rn is the same thing as doing it in two steps: first pulling back via –, and then via „.
–
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 43
at the direction of the tangent vector at (0, 0), which corresponds to t = 0 for all three
parametrizations. We have –Õ (0) = (1, 0), — Õ (0) = (1, 0), and “ Õ (0) = (≠1, 0). So – and —
induce the same orientation on the curve (from (0, 0) to (1, 1)), while “ induces the opposite
orientation. Therefore, we expect the integral of Ê over – and — to both give the same answer,
while the integral over “ should pick a sign.
Let us do the calculation for fun. The integral over – was already performed in Exam-
ple 3.3.3. For —, we get:
⁄ ⁄
Ê= —úÊ
— [0,ln(2)]
⁄ ln(2) 3 4
dx dy
= f (x(t), y(t)) + g(x(t), y(t)) dt
0 dt dt
⁄ ln(2) 1 2
= (et ≠ 1)2 et + 2et (et ≠ 1) cos(et ≠ 1) dt
0
⁄ 11 2
= u2 + 2u cos(u) du,
0
where we did the substitution u = et ≠ 1. But this is the same integral as in Example 3.3.3,
so we get the same result indeed.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 44
As for “, we get:
⁄ ⁄
Ê= “úÊ
“ [≠1,0]
⁄ 0 3 4
dx dy
= f (x(t), y(t)) + g(x(t), y(t)) dt
≠1 dt dt
⁄ 0 1 2
= ≠t2 + 2t cos(≠t) dt
≠1
⁄ 01 2
= u2 + 2u cos(u) du,
1
where we did the substitution u = ≠t. By exchanging the limits of integration, we see that
this is minus the integral in Example 3.3.3, and thus we get a minus sign as expected. ⇤
dr
T= ,
dt
2
Note that the symbol T(t) is sometimes used to denote the normalized or unit tangent vector (i.e. our
tangent vector divided by its norm |T(t)|), in which case dt should be replaced by ds = |T(t)| dt.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 45
standing for the velocity of the object moving along the curve, and the notation
dr
dr = dt
dt
is used. With this notation, one can rewrite the line integral of a vector field along the curve
as ⁄
F · dr.
C
The notation makes sense, as we know that the integral is invariant under orientation-
preserving reparametrizations, so we can rewrite it in terms of the image curve C itself, with
the orientation specified by the direction of travel along the curve. With this notation however
one needs to keep in mind that to evaluate the integral we need to compose the vector field
with the parametrization to rewrite F as a function of the parameter t before integrating in t
over the interval [a, b].
3.3.5 Exercises
1. Consider the one-form Ê = xy dx + z 2 dy + z dz on R3 . Find its pullback along the
helix centered around the z-axis and with tangent vector T(t) = (≠3 sin t, 3 cos t, 4),
0 Æ t Æ 2fi, and initial position r(0) = (3, 0, 1).
Solution. First, we find a parametrization for the helix. Integrating the tangent
vector, we know that the parametrization must be given by – : [0, 2fi] æ R3 with
–(t) = (3 cos t + A, 3 sin t + B, 4t + C) for some constants A, B, C. Using the fact that
–(0) = (3, 0, 1), we conclude that A = 0, B = 0, and C = 1. Thus the parametrization is
–(t) = (3 cos t, 3 sin t, 4t + 1).
We can then calculate the pullback –ú Ê. We get:
1 2
–ú Ê = (3 cos t)(3 sin t)(≠3 sin t) + (4t + 1)2 (3 cos t) + (4t + 1)(4) dt
1 2
= ≠27 cos t sin2 t + 3(4t + 1)2 cos t + 4(4t + 1) dt.
2. Consider the one-form Ê = x2 +y
x
2 dx + x2 +y 2 dy + z dz. Explain why you cannot pull it
y
The problem is that this is not defined at t = 1, which is part of the interval [0, 2] for the
parametric curve. In other words, the result of the pullback is not actually a well defined
one-form on an open subset containing [0, 2], since it is not defined at t = 1.
3. Find the integral of the one-form Ê = x dx + x dy + y dz in R3 along one turn clockwise
around the circle of radius two in the xy-plane and centered at the origin.
Solution. The circle has equation x2 + y 2 = 4. We parametrize one turn clockwise
around the circle with – : [0, 2fi] æ R3 , –(t) = (2 sin t, 2 cos t, 0) (recall that we are in
R3 ). The pullback of the one-form is
–ú Ê = ((2 sin t)(2 cos t) + (2 sin t)(≠2 sin t) + (2 cos t)(0)) dt = 2 (sin(2t) + cos(2t) ≠ 1) dt.
5. Let C be the curve from (0, 0, 0) to (1, 1, 1) along the intersection of the surfaces y = x2
and z = x3 . Find the integral of the vector field F(x, y, z) = (x2 , xy, z 2 ) along this curve.
Solution. We first need to parametrize the curve. A point in R3 on the surface y = x2
has coordinates (t, t2 , z) for (z, t) œ R2 . At the intersection with the surface z = x3 , we
must also have z = t3 . It then follows that points on the intersection of the two surfaces
have coordinates (t, t2 , t3 ) with t œ R. Now we want our curve to start at (0, 0, 0) and
end at (1, 1, 1). So our parameter must go from t = 0 to t = 1. We thus end up with the
parametrization
– : [0, 1] æ R3 , –(t) = (t, t2 , t3 ).
The tangent vector is T(t) = (1, 2t, 3t2 ). The line integral becomes:
⁄ ⁄ 11 2
F · T dt = (t2 , t3 , t6 ) · (1, 2t, 3t2 ) dt
C 0
⁄ 11 2
= t2 + 2t4 + 3t8 dt
0
1 2 3
= + +
3 5 9
16
= .
15
Objectives
You should be able to:
• State the Fundamental Theorem of line integrals for line integrals of exact one-forms,
and use it to evaluate line integrals.
• Show that the Fundamental Theorem of line integrals implies that line integrals of exact
one-forms only depend on the starting and ending points of the curve.
• Show that the Fundamental Theorem of line integrals implies that line integrals of exact
one-forms over closed curves vanish.
• State these results in terms of conservative vector fields and their associated potential
functions.
• Use the Fundamental Theorem of line integrals and its consequences to show that a
given one-form cannot be exact.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 48
The integral thus only depends on the starting and ending points of the image curve C.
Proof. You have probably noticed that this theorem is similar in flavour to the Fundamental
Theorem of Calculus for definite integrals; in fact it follows from it, as we will see.
First, by the definition of line integrals, we have:
⁄ ⁄
df = –ú (df ).
– [a,b]
Next, we can use one of the fundamental properties of the pullback, which is that –ú (df ) =
d(–ú f ). So we can write: ⁄ ⁄
df = d(–ú f ).
– [a,b]
If we introduce a parameter t for the parametric curve, i.e. –(t) = (x1 (t), . . . , xn (t)), then
–ú f (t) = f (–(t)), and we can write the integral as:
⁄ ⁄ b
d
df = (f (–(t)) dt.
– a dt
But then, the right-hand-side is just a standard definite integral of the derivative of a function.
By the Fundamental Theorem of Calculus (part 2), we know that the right-hand-side is simply
equal to f (–(b)) ≠ f (–(a)). We thus get:
⁄
df = f (–(b)) ≠ f (–(a)).
–
⌅
This result makes it very easy to evaluate line integrals for exact one-forms. But it also
has deeper implications. Since the integral only depends on the starting and ending points on
the image curve, this means that it does not actually depend on the choice of curve itself!
Pick any two parametric curves whose images start and end at the same place: the integral
will be the same. This is rather striking!
Corollary 3.4.2 The line integrals of an exact form along two curves that start
and end at the same points are equal. If – and — are two parametric curves whose
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 49
image share the same starting and ending points, and Ê = df is exact, then:
⁄ ⁄
Ê= Ê.
– —
Another direct consequence of the Fundamental Theorem of line integrals is that the
integral of an exact one-form over a closed curve always vanishes! Indeed, the curve is closed if
–(b) = –(a) (so that the image curve is a “loop”), and so the right-hand-side in Theorem 3.4.1
vanishes.
Corollary 3.4.3 The line integral of an exact one-form along a closed curve
vanishes. Let Ê = df be an exact one-form on an open subset U ™ Rn , and – be any closed
parametric curve whose image C µ U . Then
⁄
Ê = 0.
–
i
We sometimes write –Ê for the line integral of a one-form along a closed parametric curve.
Example 3.4.4 An example of a line integral of an exact one-form. Suppose that
you want to integrate the one-form Ê = y 2 z dx + 2xyz dy + xy 2 dz over the line segment
joining the origin to the point (1, 1, 1) œ R3 . In principle, to evaluate the line integral, you
would need to find a parametrization for the line, and use the definition of line integrals
Definition 3.3.2 to evaluate the integral. However, we notice here that Ê is exact! Indeed, if
you pick the function f (x, y, z) = xy 2 z, then
ˆf ˆf ˆf
df = dx + dy + dz = y 2 z dx + 2xyz dy + xy 2 dz,
ˆx ˆy ˆz
which is Ê. Thus we can use the Fundamental Theorem of line integrals to evaluate the line
integral. Let – be any parametrization of the line segment joining the origin to the point
(1, 1, 1). We get: ⁄
Ê = f (1, 1, 1) ≠ f (0, 0, 0) = 1 ≠ 0 = 1.
–
What is neat as well is that you know that the line integral of Ê along any curve joining
the origin to the point (1, 1, 1) will be equal to 1! The curve does not have to be a line. It
could be a parabola, the arc of a circle, anything! For instance, just for fun let us pick the
following parametric curve — : [0, 1] æ R3 with —(t) = (t, t2 , t3 ), whose image curve joins the
origin to (1, 1, 1). Let us show that this works. By definition of line integrals,
⁄ ⁄ 11 2
Ê= (t2 )2 t3 (1) + 2(t)(t2 )(t3 )(2t) + t(t2 )2 (3t2 ) dt
— 0
⁄ 11 2
= t7 + 4t7 + 3t7 dt
0
⁄ 1
=8 t7 dt
0
-1
8-
=t -
0
=1.
Neat!
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 50
One thing that we did not explain however here: how did we know that Ê was exact? This
is not always so easy to figure out. We will discuss this further in Section 3.6. ⇤
where P0 = r(a) and P1 = r(b) are the starting and ending points respectively of the curve C
with direction of travel specified by the position function r.
Note that from Corollary 3.4.2 and Corollary 3.4.3, we know that:
• the line integral of a conservative vector field does not depend on the path chosen
between two points;
• the line integral of a conservative vector field along a closed curve is always zero.
3.4.3 Exercises
1. Consider the one-form Ê = (y + zex ) dx + (x + ey sin z) dy + (z + ex + ey cos z) dz on R3 .
Show that Ê is an exact form, and use this fact to evaluate the integral of Ê along the
parametric curve – : [0, fi] æ R3 with –(t) = (t, et , sin t).
Solution. To show that it is exact, we simply find a function f (x, y, z) such that Ê = df
by inspection (we can also do that by integrating the partial derivatives as we did a
number of times already). We guess that f (x, y, z) = xy + zex + ey sin z + 12 z 2 , and check
that it works. Its differential is:
which is indeed Ê. So our guess is correct, and we have shown that Ê is exact.
Using the Fundamental Theorem of line integrals, we can integrate Ê directly along
–:
⁄
Ê =f (–(fi)) ≠ f (–(0))
–
=f (fi, efi , 0) ≠ f (0, 1, 0)
=fiefi .
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 51
2. Recall from Example 2.2.13 (see also Example 3.6.5) that the one-form Ê = ≠ x2 +y
y
2 dx +
x 2
dy on R \ {(0, 0)} is closed. However, we said that it was not exact. Use the
x2 +y 2
integral of Ê along one turn counterclockwise around the unit circle to show that Ê
cannot be exact.
Solution. We parametrize the unit circle as usual by – : [0, 2fi] æ R2 with –(t) =
(cos t, sin t). The pullback one-form is
3 4
sin t cos t
– Ê= ≠ 2
ú
2 (≠ sin t) + (cos t) dt
cos t + sin t cos t + sin2 t
2
=(sin2 t + cos2 t) dt
=dt.
In particular, it is non-zero. This proves that Ê cannot be exact on R2 \ {(0, 0)}, since if
it was exact its line integral along a closed curve would have to vanish.
3. Suppose that F is a conservative vector field in R2 and that its integral from point (1, 0)
to (≠1, 0) along the upper half of the unit circle is 5. What should the integral from
(1, 0) to (≠1, 0) but along the lower half of the circle be?
Solution. It should be 5! Indeed, since F is conservative, we know that its line integral
does not depend on the path chosen between two points. Since both paths here start
and end at the same points, the line integrals along these paths must be equal.
4. Consider the one-form Ê = df on R2 with f (x, y) = sin(x + y). Find a parametric curve
– that is not closed but such that ⁄
Ê = 0.
–
But we don’t want a closed curve, so we must choose our curve such that –(b) ”= –(a).
There are of course many possible choices. Here is one example:
Then
sin(x(a) + y(a)) = sin(0) = 0, sin(x(b) + y(b)) = sin(fi) = 0.
Thus ⁄
Ê=0
–
for any parametric curve that starts at (0, 0) and ends at (fi, 0). For instance, we could
pick a straight line between the two points.
5. Let Ê be a one-form that is defined on all of R2 . Let P0 , P1 , P0Õ , P1Õ be any four points in
R2 . Suppose that ⁄ ⁄
Ê= Ê
C1 C2
for any two curves C1 and C2 that start at P0 and end at P1 . Show that it implies that
⁄ ⁄
Ê= Ê
C1Õ C2Õ
for any two curves C1Õ and C2Õ that start at P0Õ and end at P1Õ .
In other words, if the line integral of a one-form between two given points is path
independent, then it is path independent everywhere.
Solution. The proof is fairly intuitive. Fix P0 , P1 œ R2 , and pick any two other points
P0Õ , P1Õ œ R2 . Let D0 be a fixed curve from P0 to P0Õ , and D1 a fixed curve from P1Õ to P1 .
Suppose that C1Õ and C2Õ are two curves from P0Õ to P1Õ .
On the one hand, the curve C1 = D0 fi C1Õ fi D1 is curve from P0 to P1 . The line
integral of Ê along C1 is
⁄ ⁄ ⁄ ⁄
Ê= Ê+ Ê+ Ê.
C1 D0 C1Õ D1
On the other hand, the curve C2 = D0 fi C2Õ fi D1 is also a curve from P0 to P1 . The line
integral of Ê along C2 is
⁄ ⁄ ⁄ ⁄
Ê= Ê+ Ê+ Ê.
C2 D0 C2Õ D1
Equating the two expressions for these line integrals, and simplifying, we end up with
the statement that ⁄ ⁄
Ê= Ê.
C1Õ C2Õ
Since this must be true for any points P0Õ and P1Õ , and any curves C1Õ and C2Õ from P0Õ to
P1Õ , we conclude that the line integral of Ê is path independent everywhere.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 53
Objectives
You should be able to:
3.5.1 Work
If you pick something off the ground, you expend energy. In physics, this is called “work”,
because you move an object that is under the influence of a force field. If you move an object
along a given path in a force field, how can you find the work done?
The idea is to use the well known “slicing” principle that turns a problem into an integration
question. Suppose that there is a force field F(x, y, z) in R3 , and that we move an object
along a path specified by a position vector r(t) with a Æ t Æ b. If the force field was constant
and did not depend on the position (x, y, z) of the object, from general physics principles
the work done on the object moving along the path would be F · (r(b) ≠ r(a)), i.e. the dot
product of the force and the displacement (in other words, it is the magnitude of the force
times the displacement in the direction of the force). However, as the force field depends on
the position of the object, we cannot easily calculate the work directly. But we can slice the
problem and sum over slices to rewrite the calculation as an integral.
We slice our time interval [a, b] into small slices of width t. Over a small time interval
t, the object moves from position r(t) to position r(t) + r, where r = ( x, y, z). If
the time interval is small, we can assume that the force is constant, and the work done during
this time interval can be calculated, to first order, by F(r(t)) · r. Now we sum over slices,
and take the limit where we have an infinite number of infinitesimal time intervals; this turns
the calculation into a definite integral:
⁄ b ⁄ b
dr
W = F · dr = F·
dt.
a a dt
We of course recognize the line integral of a vector field as defined in Lemma 3.3.7. This is
why line integrals are called work integrals: if the vector field is a force field, the line integral
over a parametrized curve calculates the work done when the objects moves along this curve.
Example 3.5.1 Work done by a (non-conservative) force field. Consider the force
field in R2 given by F(x, y) = (y, 5x). We will calculate the work done when moving an object
along two closed curves:
1. going once around the unit circle counterclockwise, starting and ending at (1, 0);
2. going once around a square counterclockwise, with vertices (1, 1), (1, ≠1), (≠1, ≠1), (≠1, 1),
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 54
d2 r
F=m ,
dt2
where m is the mass of the object. From the discussion above, we see that the work done by
the force on the object is:
⁄
W = F · dr
–
⁄ b 2
d r dr
=m · dt
a dt2 dt
⁄ b 3 4 ⁄ b
d dr dr dr d2 r
=m · dt ≠ m · 2 dt.
a dt dt dt a dt dt
We notice that the second integral on the last line is the same as the integral on the previous
line, so we end up with the statement that
⁄ b 2 ⁄ 3 4
d r dr 1 b d dr dr
m · dt = m · dt
a dt2 dt 2 a dt dt dt
m m
= |rÕ (b)|2 ≠ |rÕ (a)|2 .
2 2
Since rÕ (t) = v(t) is the velocity of the object, we recognize the expression m 2
2 |r (t)| as being
Õ
the kinetic energy of the object at the point parametrized by t on its path C, which we will
denote by T (t). Thus we conclude that the work is the difference in kinetic energy
W = T (b) ≠ T (a)
between the kinetic energy T (b) of the object at the ending point of the path and its kinetic
energy T (a) at the starting point.
Now assume that F is a conservative force field. It can then be written as the gradient
of a potential. We now change our conventions (only for this section), to be consistent with
the physics literature, and introduce a minus sign. We write: F = ≠ÒV for some potential
function V . By the Fundamental Theorem of line integrals, the work can then be evaluated
as follows:
⁄
W = F · dr
–
⁄ b
=≠ ÒV · dr
a
= ≠ (V (r(b)) ≠ V (r(a)).
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 56
3.5.3 Exercises
1. The force exerted by an electric charge at the origin on a charged particle at a point
(x, y, z) œ R3 is
K
F(x, y, z) = 2 (x, y, z),
(x + y + z 2 )3/2
2
where K is a constant. Find the work done as the particle moves along a straight line
from (1, 0, 0) to (2, 2, 3).
Solution. We parametrize the line as – : [0, 1] æ R3 with –(t) = (1 + t, 2t, 3t). The
tangent vector is T(t) = (1, 2, 3). The line integral thus becomes
⁄ ⁄ 1
1
F · dr = K ((1 + t) + 2(2t) + 3(3t)) dt.
– 0 ((1 + t)2 + (2t)2 + (3t)2 )3/2
We do the substitution u = (1+t)2 +(2t)2 +(3t)2 , with du = 2 ((1 + t) + 2(2t) + 3(3t)) dt,
and u(0) = 1, u(1) = 4 + 4 + 9 = 17. The integral becomes
⁄ ⁄ 17
K
F · dr = u≠3/2 du
– 2 1
3 4
1
=K 1 ≠ Ô .
17
2. True or False. A force field F(x, y, z) = k(x, y, z), with k any constant, does no work on
a particle that moves once around the unit circle in the xy-plane.
Solution. This is true, since the force field is conservative, and the integral of a
conservative vector field around a closed curve is always zero. To show that the force
field is conservative, consider the potential f (x, y, z) = k2 (x2 + y 2 + z 2 ). Then
Òf = k(x, y, z) = F.
While the above is a sufficient solution, let us compute the line integral for fun, to
see that we get zero indeed. We parametrize the circle as – : [0, 2fi] æ R3 , –(t) =
(cos t, sin t, 0), with tangent vector T(t) = (≠ sin t, cos t, 0). The line integral becomes
⁄ ⁄ 2fi
F · dr =k (≠ sin t cos t + sin t cos t + 0) dt
– 0
=0,
as expected.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 57
3. Find the work done by the force field F(x, y, z) = (x2 + y, x + y, 0) when moving an
object from (1, 1, 1) to (0, 0, 0).
Solution. The question does not specify the path taken between (1, 1, 1) and (0, 0, 0);
so we can only calculate the work if the force is conservative, in which case its line integral
does not depend on the path chosen.
3 2
Fortunately, the force is conservative. Pick the potential f (x, y, z) = x3 + xy + y2 .
Then 1 2
Òf = x2 + y, x + y, 0 = F.
Then, using the Fundamental Theorem for line integrals, we calculate the work:
⁄ ⁄
F · dr = Òf · dr
C C
=f (0, 0, 0) ≠ f (1, 1, 1)
1 1
=≠ ≠1≠
3 2
11
=≠ .
6
Objectives
You should be able to:
cooordinate functions and show that they satisfy the requirements in Definition 2.2.9 and
Definition 2.2.14.)
It turns out that the answer to the question is fairly subtle. There is one simple case
however when the statement is always true: it is when the one-form is defined (and smooth,
by definition) on all of Rn .
Theorem 3.6.1 Poincare’s lemma, version I. Let Ê be a one-form defined on all of Rn .
Then Ê is exact if and only if Ê is closed.1
The corresponding statement in terms of the associated vector field F is that, if F is defined
and has continuous partial derivatives on all of R2 or R3 , then F is conservative if and only
if it passes the screening test from Lemma 2.2.11 and Lemma 2.2.16. In the language that we
will introduce in Section 4.4, in the case of R3 one can say that F is conservative if and only
if it is curl-free:
Ò ◊ F = 0.
Proof. The proof is rather interesting, and in fact constructive, as it provides a way of
calculating the function f such that Ê = df if Ê is closed. We will write the proof only for R2 ,
but a similar proof works in Rn .
First, we notice that one direction of implication is clear: we already know from Lemma 2.2.10
that exact one-forms are closed. So all we need to show is the other direction of implication,
namely that closed one-forms are exact.
Assume that Ê = f (x, y) dx + g(x, y) dy is defined on all of R2 , and that it is closed. From
Definition 2.2.9, this means that ˆf ˆy = ˆx .
ˆg
Let us now construct a function q as follows. We take our one-form Ê, and we integrate it
along a curve in R2 which consists of, first, a horizontal line from the origin (0, 0) to the point
(x0 , 0) for some fixed x0 > 0, and then a vertical line from the point (x0 , 0) to the point (x0 , y0 )
for some fixed y0 > 0. This is a piecewise-parametric curve, but we can easily parametrize
each line segment. For the line segment from (0, 0) to (x0 , 0), we use the parametrization
–1 : [0, x0 ] æ R2 with –1 (t) = (t, 0), and for the second line segment from (x0 , 0) to (x0 , y0 ),
we use the parametrization –2 : [0, y0 ] æ R2 with –2 (t) = (x0 , t). The pullbacks of the
one-form Ê are –1ú Ê = f (t, 0) dt and –2ú Ê = g(x0 , t) dt. We construct our new function as q as
the line integral of Ê along this curve:
⁄ ⁄ ⁄ x0 ⁄ y0
q(x0 , y0 ) = Ê+ Ê= f (t, 0) dt + g(x0 , t) dt.
–1 –2 0 0
q is a function of (x0 , y0 ). Now we rename the variables (x0 , y0 ) to be (x, y), and extend the
function to all (x, y) œ R2 , not just positive numbers, as the integrals on the right-hand-side
remain well defined. So we get the function
⁄ x ⁄ y
q(x, y) = f (t, 0) dt + g(x, t) dt
0 0
defined on R2 .
Our claim is that this new function q(x, y) is in fact the potential function for Ê, i.e.,
Ê = dq, which would of course show that Ê is exact. So let us compute dq. To do so, we need
ˆx and ˆy . First,
ˆq ˆq
⁄ x ⁄ y
ˆq ˆ ˆ
= f (t, 0) dt + g(x, t) dt
ˆy ˆy 0 ˆy 0
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 59
=g(x, y)
where we used the Fundamental Theorem of Calculus part 1 for the second integral (recalling
that x is kept fixed when we evaluate the partial derivative with respect to y) and the fact
that the first integral does not depend on y at all. As for the partial derivative with respect
to x, we get:
⁄ x ⁄
y
ˆq ˆ ˆ
= f (t, 0) dt + g(x, t) dt
ˆx ˆx 0 ˆx 0
⁄ y
ˆg(x, t)
=f (x, 0) + dt by FTC Part 1 for the first term
0 ˆx
⁄ y
ˆf (x, t) ˆg(x, t) ˆf (x, t)
=f (x, 0) + dt since = , as Ê is closed,
0 ˆt ˆx ˆt
=f (x, 0) + f (x, y) ≠ f (x, 0)
=f (x, y).
Therefore,
ˆq ˆq
dq = dx + dy = f (x, y) dx + g(x, y) dy = Ê,
ˆx ˆy
and we have shown that Ê is exact. Moreover, we found an explicit expression for the potential
function as a line integral of Ê. ⌅
So we now have a clear criterion to determine whether a one-form on R is exact: we
n
simply need to show that is closed. In terms of vector fields, all we have to do is show that it
passes the screening test.
Example 3.6.2 Closed forms are exact. Consider the example from Example 3.4.4. The
one-form was Ê = y 2 z dx + 2xyz dy + xy 2 dz, and we know that it is exact, as it can be
written as Ê = df for f (x, y, z) = xy 2 z. But suppose that we don’t know that. How can we
determine quickly whether it is exact or not?
First, we notice that Ê is well defined on all of R3 . So to determine that it is exact, all
that we need to do is show that it is closed.
Let us write Ê = f1 dx + f2 dy + f3 dz. We calculate partial derivatives:
ˆf1 ˆf1 ˆf2
= 2yz, = y2, = 2xy,
ˆy ˆz ˆz
and
ˆf2 ˆf3 ˆf3
= 2yz, = y2, = 2xy.
ˆx ˆx ˆy
The statement that Ê is closed is just that the partial derivatives in the first line are equal
to the partial derivatives in the second line, which is indeed true. Thus Ê is closed, and by
Poincare’s lemma we can conclude that it must be exact.
That doesn’t tell us how to find the potential function f though. To find f , we proceed as
usual. Let us do it here for completeness.
1
To be precise, we only defined closeness for one-forms in R2 and R3 at this stage. But the statement holds
in Rn using the general definition of closed one-forms in Definition 4.6.1 -- see Theorem 4.6.4.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 60
ˆf
= y 2 z.
ˆx
We can integrate the partial derivative -- the “constant of integration” here will be any function
g(y, z) that is independent of x. Thus we get:
⁄ x
f= y 2 z dx + g(y, z) = xy 2 z + g(y, z).
0
Next, we want
ˆf
= 2xyz.
ˆy
Using the fact that f = xy 2 z + g(y, z), this equation reads
ˆg ˆg
2xyz + = 2xyz … = 0.
ˆy ˆy
Integrating, we get:
g(y, z) = h(z),
where h(z) is a function of z alone. Putting this together, we have f = xy 2 z + h(z). Finally,
we must satisfy the remaining equation:
ˆf
= xy 2 .
ˆz
Using the fact that f = xy 2 z + h(z), this becomes
dh dh
xy 2 + = xy 2 … =0 … h = C,
dz dz
for some constant C. As we are only interested in one function f such that df = Ê, we can set
the constant C = 0. We obtain that Ê = df with f (x, y, z) = xy 2 z, as stated in Example 3.4.4.
⇤
In fact, we can go a little further and state the following theorem, which gives equivalent
formulations of what it means for a one-form to be exact (or a vector field to be conservative)
on all of Rn .
Theorem 3.6.3 Equivalent formulations of exactness on Rn . Let Ê be a one-form
defined on all of Rn , and F its associated vector field. Then the following statements are
equivalent:
1. Ê is exact (F is conservative).
In other words, if one of these statements is true, then all four statements are true.
Proof. We want to prove equivalence of the four statements. To do so, it is sufficient to prove
that (1) ∆ (2), (2) ∆ (3), (3) ∆ (4), and (4) ∆ (1). We will write a proof only for R2 .
(1) ∆ (2). All exact one-forms are closed, see Lemma 2.2.10.
(2) ∆ (3). By Poincare’s lemma, Theorem 3.6.1, we know that closed one-forms defined on
all of Rn are exact, so (1) … (2). We also know that if Ê is exact, then its line integral along
closed curves always vanishes: this is Corollary 3.4.3, which follows from the Fundamental
Theorem of line integrals. So (2) ∆ (3).
(3) ∆ (4). This follows from Exercise 3.4.3.5. Indeed, suppose that P0 is on the closed
curve that you are integrating along. Pick P1 = P0 . Then we know that the line integral of
Ê is path independent for all curves starting at P0 and ending at P1 = P0 , since by (3) the
line integrals all vanish. It then follows from Exercise 3.4.3.5 that the line integrals are path
independent everywhere, which is (4).
(4) ∆ (1). For this one we need to do a bit more work. We want to show that if the line
integrals of Ê = f dx + g dy are path independent, then Ê is exact. We proceed like in the
proof of Theorem 3.6.1. First, consider a curve C1 which consists in a horizontal line from
(0, 0) to a fixed point (x0 , 0), and then a vertical line from (x0 , 0) to a fixed point (x0 , y0 ), with
x0 , y0 > 0. We parametrize it by –1 : [0, x0 ] æ R2 with –1 (t) = (t, 0), and –2 : [0, y0 ] æ R2
with –2 (t) = (x0 , t). The pullbacks are –1ú Ê = f (t, 0) dt, –2ú Ê = g(x0 , t) dt. The line integral
then reads ⁄ ⁄ x0 ⁄ y0
q(x0 , y0 ) := Ê= f (t, 0) dt + g(x0 , t) dt.
C1 0 0
Next, we consider a second curve C2 which consists in a vertical line from (0, 0) to (0, y0 ),
and then a horizontal line from (0, y0 ) to (x0 , y0 ). A parametrization is —1 : [0, y0 ] æ R2 with
—1 (t) = (0, t), and —2 : [0, x0 ] æ R2 with —2 (t) = (t, y0 ). The pullbacks are —1ú Ê = g(0, t) dt,
and —2ú Ê = f (t, y0 ) dt. The line integral reads
⁄ ⁄ y0 ⁄ x0
p(x0 , y0 ) := Ê= g(0, t) dt + f (t, y) dt.
C2 0 0
Note that the two curves C1 and C2 start at (0, 0) and end at the same point (x0 , y0 ). Since
by (4) we know that the line integrals are path independent, we know that
q(x0 , y0 ) = p(x0 , y0 ).
As in the proof of Theorem 3.6.1, we then rename the variables (x0 , y0 ) æ (x, y) and extend
the domain of definition of the function q(x, y) = p(x, y) to all (x, y) œ R2 , since the integrals
on the right-hand-side remain well defined.
Since q(x, y) = p(x, y), the partial derivatives of q and p are also equal. In particular,
⁄ y
ˆq ˆ
= g(x, t) dt = g(x, y),
ˆy ˆy 0
and x ⁄
ˆq ˆp ˆ
= = f (t, y) dt = f (x, y),
ˆx ˆx ˆx 0
where in both cases we used FTC part 1. We conclude that
ˆq ˆq
dq = dx + dy = f (x, y) dx + g(x, y) dy = Ê,
ˆx ˆy
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 62
3.6.3 Exercises
1. Determine whether the one-form Ê = y 2 exy dx + (1 + xy)exy dy is exact. If it is, find a
function f such that Ê = df .
Solution. First, we notice that the component functions are smooth on R2 , so we know
that Ê is exact if and only if it is closed. We calculate the partial derivatives of the
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 63
component functions:
ˆ 2 xy ˆ
(y e ) = 2yexy + xy 2 exy , ((1 + xy)exy ) = yexy + y(1 + xy)exy .
ˆy ˆx
We see that the two expressions are equal. Thus Ê is closed, and hence it is also exact.
We are looking for a function f such that
ˆf ˆf
df = dx + dy = y 2 exy dx + (1 + xy)exy dy.
ˆx ˆy
Integrating the partial derivative in x, we get
f = yexy + g(y).
from which we conclude that g Õ (y) = 0, i.e. g(y) = C, which we set to zero. We thus
have found a function f such that Ê = df :
f (x, y) = yexy .
2. Determine whether the field F(x, y) = (ex + ey , xey + x) is conservative. If it is, find a
potential function.
Solution. The component functions are smooth on R2 , so the vector field is conservative
if and only if it passes the screening test. We calculate the partial derivatives:
ˆ x ˆ
(e + ey ) = ey , (xey + x) = ey + 1.
ˆy ˆx
As these two expressions are not equal, we conclude that the vector field is not conservative
on R2 .
3. Determine whether or not the following sets are (a) open, (b) path connected, and (c)
simply connected:
(a) S = {(x, y) œ R2 | y Ø 0}
(c) T = {(x, y) œ R2 | y ”= 0}
Solution. Recall that a set is open if for all points in the set, there is an open ball
centered at that point that lies within the set. It is path connected if any two points in
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 64
the set can be connected by a path. It is simply connected if it is path connected, and
all closed curves can be contracted to a point within the set.
(a) S consists in the upper half of the xy-plane, including the x-axis. First, it is not
open, since any point on the x-axis cannot be the centre of an open disk within S
(as points below the x-axis are not in S). It is however path connected, as any two
points can be connected by a path, and it is simply connected, as all closed curves
can be contracted to a point within S.
(b) U is the xy-plane minus the point (1, 1). It is certainly open and path connected,
but it is not simply connected as any closed curve surrounding (1, 1) cannot be
contracted to a point in U (as there is a hole at (1, 1)).
(c) T is the xy-plane with the x-axis removed. It is an open set. However, it is not
path connected, since two points on both sides of the x-axis cannot be connected
by a path within T . It then follows that it is also not simply connected.
(d) The unit circle in R2 , i.e. the solutions to the equation x2 + y 2 = 1, is not open
in R2 . Indeed, there is no point on the unit circle that can be surrounded by an
open disk within the circle itself (note that we are considering only the circle itself
here, not the disk). It is path connected, as you can connect any two points on the
circle by a path on the circle (the arc between the two points). But it is not simply
connected, as the circle itself (which is a closed loop) cannot be contracted to a
point (recall that the interior of the circle is not part of our set).
(e) The unit sphere in R3 is not open, just as for the circle in R2 . It is path connected,
and in this case it is also simply connected, as all closed loops on the sphere can be
contracted to a point within the sphere.
(f) V is R3 minus the origin. It is certainly open, as we are just removing one point
from R3 , and it is path connected, as any two points can be connected by a path
within the set V . Is it also simply connected? The answer is yes, it is simply
connected. Indeed, pick any closed curve within V ; you can always contract it
to a point in V . The hole at the origin does not create any issue here, because
we are in R3 ; informally, the point is that even if you pick, say, a closed curve in
the xy-plane that surrounds the origin, then you can still contract it to a point
within the set V , since you can move up in the z-direction while you contract (you
don’t have to stick to the xy-plane in the contraction process, as V is a subset in
R3 ). The upshot here is that it’s important to keep in mind that the interpretation
of simply-connectedness as meaning “no holes” is only true in R2 . For instance,
in R3 , you can convince yourself that the requirement that all closed curves can
be contracted to a point within the set could instead be interpreted as meaning
that there are no “missing lines” in the set. Ultimately, it is easier to just use the
definition of simply connectedness, i.e. that the set is path connected and that all
closed curves can be contracted to a point within the set, to check whether a set is
simply connected.
CHAPTER 3. INTEGRATING ONE-FORMS: LINE INTEGRALS 65
2
4. Consider the one-form Ê = 2 |y|
x
dx ≠ xy2 dy on the open subset U = {(x, y) œ R2 | y < 0}.
Determine whether Ê is exact, and if it is find a function f such that Ê = df .
Solution. Since we are restricting to y < 0, we can replace |y| = ≠y. The one-form is
then
x x2
Ê = ≠2 dx ≠ 2 dy.
y y
Since U is simply connected, Poincare’s lemma applies. We calculate partial derivatives:
3 4 A B
ˆ x x ˆ x2 x
≠2 = 2 2, ≠ 2 = ≠2 .
ˆy y y ˆx y y2
As the two expressions are not equal, we conclude that Ê is not closed on U , and thus it
cannot be exact.
We remark here that the choice of U was very important. If we had instead defined
the one-form on the open subset V = {(x, y) œ R2 | y > 0}, then we would have obtained
a different result! Indeed, on V , |y| = y. It then follows that Ê is closed, and since
Poincare’s lemma applies as V is simply connected, it is also exact on V . Indeed, one
2
can check that Ê = df with f (x, y) = xy on V .
In fact, the largest domain of definition for Ê would be U fi V , i.e. the set {(x, y) œ
R2 | y ”= 0}. However, this is not a simply connected set, so we cannot apply Poincare’s
lemma. In any case, Ê is not exact on this larger set, as it cannot be written as Ê = df
on all of U fi V .
Chapter 4
k-forms
In this section we go beyond one-forms and introduce the general theory of differential forms,
with a focus on R3 .
Objectives
You should be able to:
• Define basic one-forms as linear maps, and basic two- and three-forms as multilinear
maps.
• Define k-forms in R3 .
66
CHAPTER 4. k-FORMS 67
Remark 4.1.2 When we are working on R3 , it is customary to write (x, y, z) for (x1 , x2 , x3 ),
and we write the basic one-forms as dx, dy, and dz (instead of dx1 , dx2 and dx3 ).
This gives a rigorous meaning of these placeholders. Using this definition, we can write a
general linear map M : R3 æ R as
M = A dx + B dy + C dz,
⌃
In the same way we can define the notion of basic three-forms.
Definition 4.1.4 Basic three-forms. The basic three-form dxi · dxj · dxk , for
i, j, k œ {1, . . . , n} is a multilinear map dxi · dxj · dxk : Rn ◊ Rn ◊ Rn æ R which takes
three vectors (u, v, w) œ Rn ◊ Rn ◊ Rn , with u = (u1 , . . . , un ) and v = (v1 , . . . , vn ), and
1
A multilinear map is a function of several variables that is linear separately in each variable.
CHAPTER 4. k-FORMS 68
⌃
We wrote the definition of basic one-, two-, and three-forms explicitly for clarity, but in
fact they are just special cases of the more general definition of basic k-forms. We now define
basic k-forms, for completeness, but don’t worry if the notation is confusing you: as we will
see, in R3 all basic k-forms with k Ø 4 automatically vanish, so the above definitions are
sufficient.
Definition 4.1.5 Basic k-forms. The basic k-form dxi1 · . . . · dxik , with i1 , . . . , ik œ
{1, . . . , n}, is a multilinear map dxi1 · . . . · dxik : (Rn )k æ R which takes k vectors
(u1 , . . . , uk ) œ (Rn )k , with uj = (uj1 , . . . , ujn ), and maps them to the following determinant:
Q R
u1i1 . . . uki1
c . .. . d.
a ..
dxi1 · . . . · dxik (u1 , . . . , uk ) = det c . .. d b
u1ik k
. . . uik
⌃
It looks like there are many possibilities here, but many of them either vanish or are related
to each other. More precisely, basic k-forms satisfy the following antisymmetry relations,
which significantly reduce the number of non-zero basic forms R3 .
Lemma 4.1.6 Antisymmetry of basic k-forms. The basic two-forms satisfy the following
properties. For any i, j œ {1, . . . , n},
In particular,
dxi · dxi = 0, i œ {1, . . . , n}.
Similarly, the basic k-forms pick a sign whenever we exchange the order of two of the dxi ’s,
and vanish if two of the dxi ’s are the same.
It follows that the only non-vanishing basic k-forms in Rn are for 1 Æ k Æ n. In particular,
in R3 we only have basic one-, two-, and three-forms. The independent basic two-forms in R3
are (using the x, y, z notation for R3 ):
dx · dy · dz.
Proof. This follows directly from the property of the determinant. If we exchange a dxi with
dxj , we exchange two rows in the matrix that we take the determinant of, and hence the
determinant picks a sign. Similarly, if two of the dxi ’s are the same, then the matrix that we
CHAPTER 4. k-FORMS 69
take the determinant of has two equal rows, and hence the determinant is zero. ⌅
Remark 4.1.7 We note here that the basic k-forms can be given a geometric interpretation.
We focus on R3 for simplicity. It’s easier to start with the basic three-form in R3 . The three
vectors u, v and w in R3 span a three-dimensional parallelepiped. Then the basic three-
form dx · dy · dz(u, v, w) calculates its oriented volume, since this is what the determinant
calculates.
As for the basic two-forms, the two vector u and v in R3 span a parallelogram. The basic
two-form dx · dy(u, v) calculates the oriented area of its projection on the xy-plane, while
the two-forms dx · dz(u, v) and dy · dz(u, v) calculate the oriented area of its projection on
the xz- and yz-planes respectively.
4.1.3 k-forms in R3
We are now ready to introduce the concept of k-forms, with k œ {0, 1, 2, 3}, in R3 , which
naturally generalizes the one-forms introduced in Definition 2.1.1.
Definition 4.1.8 k-forms in R3 . Let U ™ R3 be an open subset.
f dx + g dy + h dz,
f dy · dz + g dz · dx + h dx · dy,
f dx · dy · dz,
⌃
Of course there is no point in defining k-forms with k Ø 4 in R3 ,
as those would necessarily
vanish, by Lemma 4.1.6. But we note that this definition of k-forms can naturally be
generalized to Rn using the general definition of basic k-forms in Definition 4.1.5.
Remark 4.1.9 Our definition of two- and three-forms involves a specific choice of basic two-
and three-forms. For instance, we used dx · dy · dz instead of dz · dx · dy. When we express
the differential forms with the choice of basic form in Definition 4.1.8, we say that the k-forms
are in standard form. We generally want to present differential forms in standard form, to
simplify things.
With this being said, you may wonder why we chose this particular choice of ordering for
CHAPTER 4. k-FORMS 70
f dy · dz + g dz · dx + h dx · dy
4.1.5 Exercises
1. Show that dz · dx · dy = dx · dy · dz.
Solution. We know that dz · dx = ≠dx · dz, and dz · dy = ≠dy · dz. So we can write
dz · dx · dy = ≠dx · dz · dy = dx · dy · dz.
CHAPTER 4. k-FORMS 71
Then the basic two-forms are obtained by pairing those two-by-two, up to anti-symmetry.
We get the basic two-forms:
dx1 · dx2 , dx1 · dx3 , dx1 · dx4 , dx2 · dx3 , dx2 · dx4 , dx3 · dx4 .
For the basic three-forms, we pair the one-forms three-by-three, without repeated factors
(otherwise they would vanish). We get:
dx1 · dx2 · dx3 , dx1 · dx2 · dx4 , dx1 · dx3 · dx4 , dx2 · dx3 · dx4 .
Finally, there is only one independent basic four-form, since there cannot be repeated
dxi factors. (There is always only one independent basic “top-form”, i.e. basic n-form in
Rn ). It reads:
dx1 · dx2 · dx3 · dx4 .
3. Write down the vector field F : R3 æ R3 associated to the two-form
Ê = xy dx · dy + xyz dx · dz + x2 dy · dz.
Solution. Before we extract the vector field we need to make sure that we write the
one-form in the correct form according to the dictionary Table 4.1.11. We have:
Ê = x2 dy · dz ≠ xyz dz · dx + xy dx · dy.
Solution. By definition of a basic two-form (and recalling that we use the standard
notation here that dx · dz = dx1 · dx3 ), we have:
A B
u v
dx · dz(u, v) = det 1 1 .
u3 v3
(a) Show that Ê(u, v, w) = 1, with u = (1, 0, 0), v = (0, 1, 0) and w = (0, 0, 1) basis
vectors in R3 .
(c) In general, show that there are three choices of ordering of the basis vectors for
which Ê evaluates to 1, and three choices for which it evaluates to ≠1.
(c) In general, it is easy to see that the following three orderings give 1:
The reason is that whenever we permute two basis vectors, we exchange two columns in
the matrix that we are taking the determinant of. But we know from properties of the
determinant that swapping two columns of matrix changes its detereminant by ≠1. So
we conclude that doing an even number of two-by-two swaps of basis vectors does not
change the determinant, while doing an odd numbers of two-by-two swaps changes the
determinant by ≠1.
FYI: in the language of group theory, the group of permutations of three objects
is called the “symmetric group” S3 , whose elements are the permutations. We call a
permutation that is a swap of two objects a “transposition”. All permutations can be
obtained by doing a finite number of transpositions (i.e. by swapping objects two-by-two
a finite number of time). We define the “sign” of a permutation as being positive if
the permutation can be obtained by an even number of transpositions, and negative if it
requires an odd number of transpositions. The statement above would then be that the
determinant is unchanged if the basis vectors are related by a positive permutation, and
picks a sign if they are related by a negative permutation.
CHAPTER 4. k-FORMS 73
Objectives
You should be able to:
• Relate the wedge product of two one-forms to the cross-product of the associated vector
fields, and the wedge product of a one-form and a two-form to the dot product of the
associated vector fields.
• State and use general properties of the wedge product for k-forms.
Ê · ÷ = (x dx + x dy + ez dz) · (y dx + x dy + z dz)
=xy dx · dx + x2 dx · dy + xz dx · dz + xy dy · dx + x2 dy · dy + xz dy · dz
+ yez dz · dx + xez dz · dy + zez dz · dz
=(xz ≠ xez )dy · dz + (≠xz + yez )dz · dx + (x2 ≠ xy) dx · dy,
• If both k and m are odd, then Ê · ÷ = ≠÷ · Ê, and the wedge product is anti-commutative.
Then
Ê · ÷ = f g (dxi1 · . . . dxik ) · (dxj1 · . . . · dxjm ) ,
while
÷ · Ê = f g (dxj1 · . . . · dxjm ) · (dxi1 · . . . dxik ) .
To relate the second expression to the first, we need to move the dxj ’s to the left of the dxi ’s.
We first move dxi1 to the left. Each time we pass a dxj , we pick a sign. So in the end we get:
We do the same thing with dxi2 , moving it pass all the dxj ’s, and so on all the way to dxik .
The final result is
Finally, the statement is proved for general differential forms by doing this manipulation term
by term after distributing the wedge product. ⌅
F = (f1 , f2 , f3 ), G = (g1 , g2 , g3 ),
what we are doing is constructing a new vector field, let’s call it H for the time being,
associated to Ê ·÷, with component functions given by (according to the dictionary established
in Table 4.1.11 for relating a two-form to a vector field):
H = (f2 g3 ≠ f3 g2 , f3 g1 ≠ f1 g3 , f1 g2 ≠ f2 g1 ) .
What is this vector field? It is nothing but the cross-product of the vector fields F and G!
Indeed,
F ◊ G = (f2 g3 ≠ f3 g2 , f3 g1 ≠ f1 g3 , f1 g2 ≠ f2 g1 ) .
Therefore, we end up with the following statement:
Lemma 4.2.7 The wedge product of two one-forms is the cross-product of the
associated vector fields. If Ê and ÷ are two one-forms on U ™ R3 , with associated vector
fields F and G, then the vector field associated to the two-form Ê · ÷ via the dictionary in
Table 4.1.11 is the cross-product F ◊ G.
Finally, we consider the wedge product of a one-form and a two-form. Let Ê = f1 dx +
f2 dy + f3 dz and ÷ = g1 dy · dz + g2 dz · dx + g3 dx · dy. Then the wedge product is the
three-form:
Ê · ÷ = (f1 g1 + f2 g2 + f3 g3 ) dx · dy · dz.
According to the dictionary in Table 4.1.11, we thus conclude that the function associated
to the three-form Ê · ÷ is nothing but the dot product of the two vector fields F and G
associated to Ê and ÷ respectively:
F · G = f1 g1 + f2 g2 + f3 g3 .
Table 4.2.10 Dictionary between the wedge product of a one-form and a two-form
in R3 and vector calculus concepts
Differential form concept Vector calculus concept
1-form Ê vector field F
2-form ÷ vector field G
3-form Ê · ÷ vector field F · G
4.2.3 Exercises
1. Let Ê = ex dx + y dz and ÷ = xy dx + z dy + y dz. Find Ê · ÷, and write your result in
standard form.
Solution. We find (using the fact that dx · dx = dy · dy = dz · dz = 0):
vector fields in Rn as follows: take two one-forms Ê and ÷ on Rn , with associated vector
fields F and G. We would define the “cross-product” of the vector fields as being given
by the two-form Ê · ÷ in Rn . Note that only on R3 can we associate to the result a new
vector field (this is because of Hodge duality between one-forms and two-forms in R3 ,
see Section 4.8). For instance, in R4 a two-form has 6 component functions (there are 6
independent non-vanishing basic two-forms in R4 ), so it cannot be associated to a vector
field.
4. Let Ê = f dx + g dy + h dz be an arbitrary one-form on R3 . By doing an explicit
calculation, show that Ê · Ê = 0.
Solution. We find:
According to the dictionary Table 4.1.11, the vector field associated to this two-form is
This is indeed the cross-product, as you can calculate using standard formulae from
linear algebra. For instance, you may have seen the formula:
Q R
i j k
c d
F ◊ G = det a F1 F2 F3 b
G 1 G2 G3
Q R
i j k
c d
= det a1 1 0 b
0 2 3
=3i ≠ 3j + 2k,
A · (B ◊ C) = B · (C ◊ A) = C · (A ◊ B).
CHAPTER 4. k-FORMS 79
Prove this property by looking at the wedge product of the three one-forms Ê, ÷, ⁄
associated to the vectors A, B, C.
Solution. As discussed in Lemma 4.2.7 and Lemma 4.2.8, we know that
Ê · ÷ · ⁄ = A · (B ◊ C)dx · dy · dz,
÷ · ⁄ · Ê = B · (C ◊ A)dx · dy · dz,
and
⁄ · Ê · ÷ = C · (A ◊ B)dx · dy · dz.
But since ⁄, Ê, ÷ are one-forms, we know that Ê · ÷ = ≠÷ · Ê, ÷ · ⁄ = ≠⁄ · ÷, and
Ê · ⁄ = ≠⁄ · Ê. Therefore,
Ê · ÷ · ⁄ = ÷ · ⁄ · Ê = ⁄ · Ê · ÷,
and the statement is proved. In other words, it follows directly from anti-commutativity
of the wedge product of one-forms.
Objectives
You should be able to:
• Define the exterior derivative of a k-form, focusing on zero-, one- and two-forms in R3 .
• State and use the graded product rule for the exterior derivative.
• Show that applying the exterior derivative twice always gives zero.
where d(fi1 ···ik ) means the exterior derivative of the zero-form (function) fi1 ···ik . In other
words, we are applying the exterior derivative d to the component functions of Ê. ⌃
This definition may seem a little daunting because of the summations, so let us be more
explicit for k-forms in R3 .
Lemma 4.3.2 The exterior derivative in R3 .
And that’s it. In particular, the exterior derivative of a three-form on R3 must vanish, since it
would give a four-form, but all k-forms with k Ø 4 necessarily vanish on R3 by Lemma 4.1.6.
Proof. We start with Definition 4.3.1 restricted to the case with n = 3. For the exterior
derivative of a zero-form, the statement is obvious. For the exterior derivatives of a one-form
and a two-form, all we have to do is evaluate the exterior derivatives d(f ), d(g) and d(h) of
the zero-form f, g, h, and rearrange terms using Lemma 4.1.6. ⌅
Note that you certainly should not aim at learning these formulae by heart. The whole
point is precisely that you don’t need to learn these formulae! All you need to remember is
CHAPTER 4. k-FORMS 81
that, to evaluate the exterior derivative of a k-form, you act with the exterior derivative on
the component functions of the k-form. This may be clearer with examples.
Example 4.3.3 The exterior derivative of a zero-form on R3 . We are already familiar
with the calculation of the exterior derivative of a zero-form, since this is the same thing as
the calculation of the differential of a function that we defined in Definition 2.2.1. But let us
give an example here for completeness.
Let f (x, y, z) = y ln(x) + z be a smooth function on U = {(x, y, z) œ R3 | x > 0}. Its
exterior derivative is the one-form df on U given by:
ˆf ˆf ˆf
df = dx + dy + dz
ˆx ˆy ˆz
y
= dx + ln(x) dy + dz.
x
⇤
Example 4.3.4 The exterior derivative of a one-form on R3 . Let Ê = xy dx + (z +
y) dy + xyz dz be a one-form on R . Its exterior derivative is the two-form dÊ on R3 given by:
3
⇤
Example 4.3.5 The exterior derivative of a two-form on R3 . Let Ê = (x2 + y 2 ) dy ·
dz + sin(z) dz · dx + cos(xy) dx · dy be a two-form on R3 . Its exterior derivative is the
three-form dÊ on R3 given by:
We see that going from the second line to the third line, most terms vanish, since anytime we
take the wedge of dx with itself we get zero, and same for dy and dz. ⇤
To end this section, we note that the exterior derivative is linear. If Ê and ÷ are two
k-forms, and a, b œ R, then
d(aÊ + b÷) = adÊ + bd÷.
This is proven in Exercise 4.3.4.3.
Now that we have defined the exterior derivative for differential forms, and that we know how
to multiply differential forms using the wedge product, we could ask whether the exterior
derivative satisfies a similar product rule with respect to the wedge product. It turns out that
it does, but with a twist (or more precisely a sign). We call this the “graded product rule” for
the exterior derivative.
Lemma 4.3.6 The graded product rule for the exterior derivative. Let Ê be a k-form
and ÷ be an l-form on U ™ Rn . Then:
d(Ê · ÷) =d(f g)
n
ÿ ˆ
= (f g) dxi
i=1
ˆxi
ÿn 3 4
ˆf ˆg
= g+f dxi
i=1
ˆxi ˆxi
=d(f )g + f d(g).
To prove the general statement we unfortunately need to use lots of summations. Let us
introduce the following notation for the k-form Ê and the l-form ÷:
ÿ ÿ
Ê= wi1 ···ik dxi1 · · · · · dxik , ÷= hj1 ···jl dxj1 · · · · · dxjl .
1Æi1 <···<ik Æn 1Æj1 <···<jl Æn
To go from the first to the second line, we used the fact that the product rule is satisfied for
the exterior derivative of the product of zero-forms, as shown above.
Now let us study the terms on the right-hand-side of the equation we just obtained. First,
we get:
ÿ ÿ
d(wi1 ···ik )hj1 ···jl · dxi1 · · · · · dxik · dxj1 · · · · · dxjl
1Æi1 <···<ik Æn 1Æj1 <···<jl Æn
Q R Q R
ÿ ÿ
=a d(wi1 ···ik ) · dxi1 · · · · · dxik b · a hj1 ···jl dxj1 · · · · · dxjl b
1Æi1 <···<ik Æn 1Æj1 <···<jl Æn
= d(Ê) · ÷
All that we did to go from the first line to the second line is move the zero-form (or function)
hj1 ···jl to the right of the dxi ’s, which we can do since it is a zero-form and hence commutes
with the dxi ’s by Lemma 4.2.6.
CHAPTER 4. k-FORMS 83
That takes care of the first set of terms. The remaining ones take the form
ÿ ÿ
wi1 ···ik d(hj1 ···jl ) · dxi1 · · · · · dxik · dxj1 · · · · · dxjl .
1Æi1 <···<ik Æn 1Æj1 <···<jl Æn
We would like to do the same and commute d(hj1 ···jl ) to the right of the dxi ’s, so that we can
identify this term as Ê · d(÷). However, d(hj1 ···jl ) is a one-form, and thus by Lemma 4.2.6
d(hj1 ···jl ) · dxi = ≠dxi · d(hj1 ···jl ). So every time we commute d(hj1 ···jl ) past a dxi , we pick a
sign. Since there are k dxi ’s in this expression, this tells us that it is equal to
(≠1)k Ê · d(÷).
We then calculate the exterior derivative, which gives the following three-form:
d(Ê) =d(xy) · dz
=ydx · dz + xdy · dz,
and
=dy · dx + dz · dx,
d(Ê · ÷) =(ydx · dz + xdy · dz) · ((y + z) dx + 2 dy) ≠ (xy dz) · (dy · dx + dz · dx)
=(≠2y + x(y + z) + xy)dx · dy · dz,
4.3.3 d2 = 0
There’s another fundamental property of the exterior derivative: if we apply the exterior
derivative twice, we always get zero. This may seem surprising, as this is certainly not true for
the ordinary derivative d/dx, but it is true for the exterior derivative because of antisymmetry
of the wedge product. More precisely:
Lemma 4.3.9 d2 = 0. Let Ê be a k-form on U ™ Rn . Then
d(d(Ê)) = 0.
In other words, if we apply the exterior derivative twice on any differential form, we always
get zero. We often abbreviate this statement as d2 = 0, meaning that applying the exterior
derivative twice always gives zero.
q
Proof. We write Ê = 1Æi1 <···<ik Æn wi1 ···ik dxi1 · · · · · dxik . We have:
Q R
ÿ
d(d(Ê)) =d a d(wi1 ···ik ) · dxi1 · · · · · dxik b
1Æi1 <···<ik Æn
CHAPTER 4. k-FORMS 85
Q R
ÿ n
ÿ ˆwi1 ···ik
=d a dx– · dxi1 · · · · · dxik b
1Æi1 <···<ik Æn –=1
ˆx–
ÿ ÿn 3 4
ˆwi1 ···ik
= d · dx– · dxi1 · · · · · dxik
1Æi1 <···<ik Æn –=1
ˆx–
Q R
ÿ n ÿ
ÿ n
ˆ 2 wi1 ···ik
= a dx— · dx– b · dxi1 · · · · · dxik .
1Æi1 <···<ik Æn –=1 —=1
ˆx— ˆx–
If we can show that the term in brackets in the last line is zero, then clearly d(d(Ê)) = 0. We
have:
n ÿ
ÿ n
ˆ 2 wi1 ···ik n ÿ
ÿ n
ˆ 2 wi1 ···ik
dx— · dx– = dx— · dx–
–=1 —=1
ˆx— ˆx– –=1 —=1
ˆx– ˆx—
ÿn ÿ n
ˆ 2 wi1 ···ik
=≠ dx– · dx— .
–=1 —=1
ˆx– ˆx—
ˆ2w ˆ2w
In the first, line, we used the fact that ˆx—i1ˆx···ik
–
= ˆx–i1ˆx
···ik
—
by the Clairaut-Schwarz theorem,
since the coefficient functions wi1 ···ik are assumed to be smooth. In the second line, we used
the fact that dx— · dx– = ≠dx– · dx— , by Lemma 4.1.6. Finally, in the summation on the
right-hand-side, we can simply rename – to be —, and — to be –, since those are indices
that are summed over, and hence we can give them the name we want. We end up with the
statement that
n ÿ
ÿ n
ˆ 2 wi1 ···ik n ÿ
ÿ n
ˆ 2 wi1 ···ik
dx— · dx– = ≠ dx— · dx– .
–=1 —=1
ˆx— ˆx– –=1 —=1
ˆx— ˆx–
But the only two-form that is equal to minus itself is the zero two-form, that is
n ÿ
ÿ n
ˆ 2 wi1 ···ik
dx— · dx– = 0.
–=1 —=1
ˆx— ˆx–
4.3.4 Exercises
1. Let Ê = xy dx + xy dy + xz dz on U = {(x, y, z) œ R3 | x ”= 0}. Find the two-form dÊ and
write your result in standard form. What is the vector field associated to dÊ?
Solution. From the definition, we find:
3 4 3 4
y z
dÊ =d(xy) · dx + d · dy + d · dz
x x
1 y 1 z
=y dx · dx + x dy · dx +dy · dy ≠ 2 dx · dy + dz · dz ≠ 2 dx · dz
3 4 x x x x
z y
= 2 dz · dx ≠ x + 2 dx · dy.
x x
CHAPTER 4. k-FORMS 86
Using the dictionary Table 4.1.11, the vector field associated to the two-form dÊ is
3 4
z y
F(x, y, z) = 0, 2
, ≠x ≠ 2 .
x x
2. Find the three-form dÊ if Ê = xyz(dy · dz + dz · dx + dx · dy), and write your result in
standard form.
Solution. We find:
Solution. First, we show that the property holds if Ê and ÷ are 0-forms. Since those
are simply functions, let us write them as f and g. Then:
n
ÿ ˆ
d(af + bg) = (af + bg) dxi
i=1
ˆxi
n 3
ÿ 4
ˆf ˆg
= a +b dxi
i=1
ˆxi ˆxi
ÿn ÿn
ˆf ˆg
=a dxi + b dxi
i=1
ˆxi i=1
ˆxi
=adf + bdg.
For the second equality we used the fact that partial derivatives are linear.
Now we can prove the general case. Suppose that Ê and ÷ are k-forms on U ™ Rn .
Let ÿ
Ê= fi1 ···ik dxi1 · · · · · dxik
1Æi1 <···<ik Æn
and ÿ
÷= gi1 ···ik dxi1 · · · · · dxik .
1Æi1 <···<ik Æn
Then
ÿ
d(aÊ + b÷) = d(afi1 ···ik + bgi1 ···ik ) · dxi1 · · · · · dxik
1Æi1 <···<ik Æn
ÿ
= (a dfi1 ···ik + b dgi1 ···ik ) · dxi1 · · · · · dxik
1Æi1 <···<ik Æn
ÿ
=a dfi1 ···ik · dxi1 · · · · · dxik
1Æi1 <···<ik Æn
CHAPTER 4. k-FORMS 87
ÿ
+b dgi1 ···ik · dxi1 · · · · · dxik
1Æi1 <···<ik Æn
=adÊ + bd÷.
In the second equality we used the calculation above that showed that linearity holds for
0-forms.
4. Let Ê = xey dx + z dy + yex dz. Show by explicit calculation that d2 Ê = 0.
Solution. Of course, we know that d2 Ê = 0 since this is true for any differential form
Ê, as proven in Lemma 4.3.9. But let us show that it is true by explicit calculation for
this particular one-form Ê.
We first calculate the two-form dÊ:
dÊ =d(xey ) · dx + dz · dy + d(yex ) · dz
=xey dy · dx + ey dx · dx + dz · dy + yex dx · dz + ex dy · dz
=(ex ≠ 1) dy · dz ≠ yex dz · dx ≠ xey dx · dy.
d2 Ê =d(dÊ)
=d(ex ≠ 1) · dy · dz ≠ d(yex ) · dz · dx ≠ d(xey ) · dx · dy
=ex dx · dy · dz ≠ yex dx · dz · dx ≠ ex dy · dz · dx ≠ xey dy · dx · dy
≠ ey dx · dx · dy
=(ex ≠ ex )dx · dy · dz
=0,
as expected.
5. Let Ê = (x2 ≠ y 2 ) dx + y dz and ÷ = (x2 + y 2 ) dy + y dz. By explicit calculation, show
that
d(Ê · ÷) = dÊ · ÷ ≠ Ê · d÷,
which is consistent with the graded product rule Lemma 4.3.6 since Ê is a one-form.
Solution. We need to calculate d(Ê · ÷), dÊ · ÷, and Ê · d÷. First, we calculate Ê · ÷:
Ê · ÷ =((x2 ≠ y 2 ) dx + y dz) · ((x2 + y 2 ) dy + y dz)
=(x4 ≠ y 4 )dx · dy + y(x2 ≠ y 2 ) dx · dz + y(x2 + y 2 ) dz · dy
= ≠ y(x2 + y 2 ) dy · dz ≠ y(x2 ≠ y 2 ) dz · dx + (x4 ≠ y 4 )dx · dy.
Then
d(Ê · ÷) = ≠ d(y(x2 + y 2 )) · dy · dz ≠ d(y(x2 ≠ y 2 )) · dz · dx + d(x4 ≠ y 4 ) · dx · dy
= ≠ 2xy dx · dy · dz ≠ (x2 ≠ 3y 2 ) dy · dz · dx
=(3y 2 ≠ 2xy ≠ x2 )dx · dy · dz.
Next, we calculate dÊ · ÷. We have:
dÊ =d(x2 ≠ y 2 ) · dx + dy · dz
CHAPTER 4. k-FORMS 88
= ≠ 2ydy · dx + dy · dz
=dy · dz + 2ydx · dy.
Then
d÷ =d(x2 + y 2 ) · dy + dy · dz
=2x dx · dy + dy · dz
=dy · dz + 2x dx · dy.
Then
Solution. First, using the graded product rule Lemma 4.3.6, we get that
d(f g) = f dg + gdf,
since f is a 0-form (here we omit the wedge product symbol, since we are either multiplying
two functions or a function with a one-form). By the Fundamental Theorem of line
integrals, we know that
⁄
d(f g) = f (–(b))g(–(b)) ≠ f (–(a))g(–(a)),
–
since d(f g) is an exact one-form. Using the graded product rule, we thus conclude that
⁄ ⁄ ⁄
d(f g) = f dg + gdf = f (–(b))g(–(b)) ≠ f (–(a))g(–(a)).
– – –
s
Solving for – f dg, we get the desired statement.
9. Let Ê and ÷ be one-forms that differ by the exterior derivative of a 0-form, that is,
÷ = Ê + df
for some function f . Show that
d(Ê · ÷) = dÊ · df.
Solution. We have:
d(Ê · ÷) = d(Ê · (Ê + df )) = d(Ê · Ê) + d(Ê · df ).
The first term on the right-hand-side is zero, since Ê · Ê = 0 for a one-form (see
Lemma 4.2.6; the wedge product is anti-commutative for odd forms). As for the second
term, we use the graded product rule and the fact that d2 = 0 to get:
d(Ê · ÷) = dÊ · df ≠ Ê · d2 f = dÊ · df.
CHAPTER 4. k-FORMS 90
Objectives
You should be able to:
• Derive various vector calculus identities from the graded product rule for the exterior
derivative and the statement that d2 = 0.
We define the curl of F, and denote it by Ò ◊ F, to be the vector field associated to the
two-form dÊ according to Table 4.1.11:
3 4
ˆf3 ˆf2 ˆf1 ˆf3 ˆf2 ˆf1
Ò◊F= ≠ , ≠ , ≠ .
ˆy ˆz ˆz ˆx ˆx ˆy
Note that input of the curl is a vector field, and the output is also a vector field. ⌃
Finally, we can apply the exterior derivative to a two-form to get a three-form.
Definition 4.4.3 The divergence of a vector field. Let ÷ = f1 dy · dz + f2 dz · dx +
f3 dx · dy be a two-form on U ™ R3 , with its associated vector field F = (f1 , f2 , f3 ). Its
exterior derivative d÷ is the three-form
3 4
ˆf1 ˆf2 ˆf3
d÷ = + + dx · dy · dz.
ˆx ˆy ˆz
We define the divergence of F, and denote it by Ò · F, to be the function associated to the
three-form d÷ according to Table 4.1.11:
ˆf1 ˆf2 ˆf3
Ò·F= + + .
ˆx ˆy ˆz
Note that the input of the diveregence is a vector field, and the output is a function. ⌃
Remark 4.4.4 Now you can start to see the power of developing the framework of differential
forms. These three operators, namely grad, div, and curl, which appear as independent
operators in vector calculus, are just the action of the same operator, namely the exterior
derivative, but on zero-, one-, and two-forms respectively. Moreover, we don’t need to
remember these definitions by heart: all we need to remember is how to act with the exterior
derivative on k-forms, which simply amounts to acting with the exterior derivative on the
component functions. So much simpler!
Even more powerful is the fact that the framework of differential forms naturally extend
to any dimension, not only R3 . However, the defintions of grad, curl, div, the cross-product,
etc. rely on the geometry of R3 . The natural generalization to higher dimensions is just the
action of the exterior derivative as we defined it.
Remark 4.4.5 The introduction of the curl of a vector field allows us to rephrase the
screening test for conservative vector fields in R3 in a nicer way. Looking at the screening test
in Lemma 2.2.16, it is clear that the screening test for a vector field F is satisfied if and only if
Ò ◊ F = 0.
In other words, the screening test was simply saying that the curl of the vector field vanishes.
Remark 4.4.6 Just as for the cross product of two vectors, in standard vector calculus
textbooks a determinant formula is usually given to remember how to calculate the curl of a
vector field F = (f1 , f2 , f3 ) in R3 :
Q R
i j k
c ˆ d
Ò ◊ F = det a ˆx
ˆ ˆ
ˆy ˆz b ,
f1 f2 f3
CHAPTER 4. k-FORMS 92
where i, j, k are the unit vectors in the x, y, z directions. You can use this formula if you
want. Or you can remember that the curl is obtained by taking the exterior derivative of the
one-form associated to F.
Example 4.4.7 Maxwell’s equations. Maxwell’s equations form the foundations of
electromagnetism. It turns out that they are written in terms of the divergence and the curl.
More precisely, if E is the electric vector field, and B is the magnetic vector field, both defined
on R3 (our space), Maxwell’s equations state that
Ò · E =4fifl,
Ò · B =0,
1 ˆB
Ò◊E+ =0,
c ˆt
1 ˆE 4fi
Ò◊B≠ = J,
c ˆt c
where c is the speed of light, fl is the total electric charge density, and J is the total electric
current density (which is a vector field). In particular, the equations simplify when there is
no charge or current (such as in vacuum), with fl = J = 0.
Note that we are abusing notation a little bit here. Those equations are the “time-
dependent” Maxwell’s equations. What this means is that we think of E and B as vector
fields in R3 (in space), but that also depend on another variable t corresponding to time. This
is why the equations above include partial derivatives of E and B with respect to t. The
“time-independent” Maxwell’s equations, in which E and B are true vector fields on R3 (with
no time dependence), would correspond to setting the two terms involving partial derivatives
with respect to t to zero. ⇤
3
We can summarize the dictionary between the exterior derivative in R and vector calculus
operations in the following table.
Table 4.4.8 Dictionary between the exterior derivative in R3 and vector calculus
concepts
Differential form concept Vector calculus concept
d of a 0-form df gradient Òf
d of a 1-form dÊ curl Ò◊F
d of a 2-form d÷ divergence Ò · F
1.
Ò(f g) = (Òf )g + f (Òg),
2.
Ò ◊ (f F) = (Òf ) ◊ F + f Ò ◊ F,
3.
Ò · (f F) = (Òf ) · F + f Ò · F,
4.
Ò · (F ◊ G) = (Ò ◊ F) · G ≠ F · (Ò ◊ G).
Proof. This is just the reformulation in terms of vector fields of the four different non-vanishing
cases of the graded product rule for differential forms in R3 (see Remark 4.3.8). ⌅
Now you may be starting to like this. Learning this kind of vector calculus identities by
heart is frustrating. But these are just reformulations of the one and only graded product
rule for the exterior derivative Lemma 4.3.6, which is all that you have to remember (sure,
there is an annoying sign in the graded product rule, but it’s much better than learning vector
calculus identities!).
1.
Ò ◊ (Òf ) = 0 (curl of grad is zero),
2.
Ò · (Ò ◊ F) = 0 (div of curl is zero).
Proof. This is just the reformulation of the statement that d2 = 0 for a zero-form and a
one-form in R3 . ⌅
The game of translating easy statements for the exterior derivative into complicated
statements for grad, curl, and div is fun, isn’t it? Let’s prove one more vector calculus identity
for now, which follows by combining the statement that d2 = 0 with the graded product rule.
Lemma 4.4.11 Vector calculus identities, part 3. Let f, g, h be smooth functions on
U ™ R3 . Then:
Ò · (f (Òg ◊ Òh)) = Òf · (Òg ◊ Òh).
CHAPTER 4. k-FORMS 94
Proof. Consider the action of the exterior derivative on the two-form f dg · dh:
where we used the graded product rule and the fact that f is a zero-form. Using the graded
product rule again, we know that
as claimed. ⌅
1.
Ò(F · G) = F ◊ (Ò ◊ G) + (Ò ◊ F) ◊ G + (G · Ò)F + (F · Ò)G,
2.
Ò ◊ (F ◊ G) = F(Ò · G) ≠ (Ò · F)G + (G · Ò)F ≠ (F · Ò)G.
ˆ ˆ ˆ
(G · Ò)F = G1 F + G2 F + G3 F.
ˆx ˆy ˆz
Finally, there are a few more vector calculus identities that involve the Laplacian operator,
which in the language of differential forms requires the introduction of the Hodge star operator.
We will come back to this in Section 4.8.
CHAPTER 4. k-FORMS 95
4.4.5 Exercises
1. Let F = (xy, yz, xz) be a vector field on R3 . Find its curl Ò ◊ F and divergence Ò · F.
Solution. The curl of the vector field is given by:
Q R
i j k
c d
Ò ◊ F = det a ˆx
ˆ ˆ
ˆy
ˆ
ˆz b
xy yz xz
= ≠ yi ≠ zj ≠ xk.
Note that we could have done these calculations using differential forms. To get the
curl, we associate to F a one-form Ê = xy dx + yz dy + xz dz and calculate its exterior
derivative:
The curl Ò ◊ F is then the vector field associated to this two-form, that is Ò ◊ F =
(≠y, ≠z, ≠x).
To calculate the divergence Ò · F, we associate to F a two-form ÷ = xy dy · dz +
yz dz · dx + xz dx · dy and calculate its exterior derivative:
Therefore Ò · F = x + y + z.
2. For the following two vector fields, find their curl and divergence:
1
(a) F(x, y, z) = 2 (x, y, 0).
x + y2
1
(b) G(x, y, z) = (≠y, x, 0).
x2+ y2
Solution. Let’s solve this one using differential forms. You can do it directly using the
formulae for curl and div as well.
(a) To find the curl, we associated a one-form Ê to F:
1
Ê= 2 (x dx + y dy).
x + y2
CHAPTER 4. k-FORMS 96
for some functions g(y, z), h(y, z). The first condition then imposes that
ˆ ˆ
h(y, z) = g(y, z).
ˆy ˆz
There are many possible choices, but the simplest would be g(y, z) = h(y, z) = 0. We
would then conlude that
F = (0, xy, ≠xz)
is a vector field such that Ò ◊ F = (0, z, y).
4. Is there a vector field F on R3 such that Ò ◊ F = (x, y + xz, z)? Justify your answer.
Solution. We know that Ò · (Ò ◊ F) = 0 for any vector field F. So if there is a
vector field F such that Ò ◊ F = (x, y + xz, z), then the divergence of the vector on the
right-hand-side (let’s call it G) must be zero. But
Ò · G = 1 + 1 + 1 = 3,
which is obviously non-zero. Therefore, we conclude that there does not exist a vector
field F such that Ò ◊ F = (x, y + xz, z).
5. Let F = (xy, y 2 , xy + z) and G = (xyz, yz, z 2 ) be smooth vector fields on R3 , and
– : [0, 2fi] æ R3 be the parametric curve given by –(t) = (sin(t), cos(t), t(t ≠ 2fi)). Show
that the line integral of the vector field
F ◊ (Ò ◊ G) + (Ò ◊ F) ◊ G + (G · Ò)F + (F · Ò)G
along – is zero.
Solution. Well, you certainly do not want to evaluate this line integral, it would be
painful!
First, we notice that the parametric curve – is closed, since
–(0) = (0, 1, 0) = –(2fi).
Next, we notice that we can use the vector calculus identity 1 from Lemma 4.4.12, which
states that
Ò(F · G) = F ◊ (Ò ◊ G) + (Ò ◊ F) ◊ G + (G · Ò)F + (F · Ò)G.
CHAPTER 4. k-FORMS 98
So the vector field that we want to evaluate the line integral of is Ò(F · G). As this is
the gradient of a function, this means that the vector field is conservative. Therefore, its
integral along any closed curve vanishes. We conclude that the integral along – is zero!
6. Suppose that you study a vector field F in a lab. You measure that
for some constants –, — that you are not able to determine experimentally. However, from
theoretical considerations you know that F must be divergence-free (i.e., its divergence is
zero). Find the values of – and —.
Solution. We know that
ˆ ˆ ˆ
Ò·F= (xz + yz + x2 y) + – (yz + x2 z) + — (xyz + y)
ˆx ˆy ˆz
=z + 2xy + –z + —xy.
Since we know that Ò · F = 0, and that – and — are constants (i.e. do not depend on
x, y, z), we conclude that we must have
– = ≠1, — = ≠2.
7. Let F(x, y, z) = (f (x), g(y), h(z)) for some smooth functions f (x), g(y), h(z) on R (note
that those are functions of a single variable), and let q(x, y, z) be an arbitrary smooth
function on R3 . Show that
Ò · (F ◊ Òq) = 0.
Solution. Let us first solve it using vector calculus identities, and then provide an
alternative but equivalent solutions using differential forms. Identity 4 of Lemma 4.4.9
states that
Ò · (F ◊ G) = (Ò ◊ F) · G ≠ F · (Ò ◊ G).
Applying this to the case at hand, we get:
Ò · (F ◊ Òq) = (Ò ◊ F) · Òq ≠ F · (Ò ◊ Òq).
We then calculate the curl of F. We get:
3 4
ˆh(z) ˆg(y) ˆf (x) ˆh(z) ˆg(y) ˆf (x)
Ò◊F= ≠ , ≠ , ≠
ˆy ˆz ˆz ˆx ˆx ˆy
=0.
Moreover, from Identity 1 of Lemma 4.4.10, we know that
Ò ◊ Òq = 0.
Therefore
Ò · (F ◊ Òq) = 0.
Let us now solve the question using differential forms. Let Ê = f (x) dx + g(y) dy +
h(z) dz be the one-form associated to F. Then Ò · (F ◊ Òq) is the function associated
to d(Ê · dq). So we want to show that
d(Ê · dq) = 0.
CHAPTER 4. k-FORMS 99
since by evaluating the differentials we only get terms involving the vanishing basic
two-forms dx · dx = dy · dy = dz · dz = 0. We thus conclude that
d(Ê · dq) = 0.
Objectives
You should be able to:
• Interpret the gradient of a vector field as giving the direction and magnitude of fastest
increase.
• Interpret the divergence of a vector field in terms of expansion and contraction of a fluid.
decreases
Ô as we move away from the origin, until it reaches sea level on the circle with radius
1000. So we can think of this altitude function as representing in a circular mountain centred
at the origin.
The level curves of H are the curves of constant elevation, i.e.
for constant elevations 0 Æ C Æ 1000. Those would correspond to the level curves of elevation
on a topographical map. In this case, they are all circles centred at the origin.
The gradient of H is 3 4
ˆH ˆH
ÒH = , = ≠2(x, y).
ˆx ˆy
For any point (x, y) on the surface, this is a vector that points towards the origin, which says
that the direction of steepest increasing slope is towards the origin, as expected from a circular
mountain centred at the origin. Moreover, we see that the magnitude of the gradient vector is
Ò
|ÒH| = 2 x2 + y 2 .
This says that the slope is steeper far away from the origin, and becomes less and less steep
as we get closer to the origin.
Here is below the contour map for this function (the graph of its level curves), and also a
3D plot (both produced using Mathematica).
Figure 4.5.2 A contour map of the function H(x, y) = ≠x2 ≠ y 2 + 1000, for ≠100 Æ x Æ 100
and ≠100 Æ y Æ 100.
CHAPTER 4. k-FORMS 101
Figure 4.5.3 A 3d plot of the function z = ≠x2 ≠ y 2 + 1000 (i.e. thinking of H(x, y) as
representing the height of the surface above sea level), for ≠100 Æ x Æ 100 and ≠100 Æ y Æ 100.
⇤
1
The “right hand rule” means that if you align the thumb of your right hand along the vector, the fingers of
your right hand will “curl” around the axis of rotation in the direction of rotation.
CHAPTER 4. k-FORMS 102
Example 4.5.4 The curl of the velocity field of a moving fluid. Suppose that a
moving fluid has velocity field given by
Sketching the vector field (see Exercise 2.1.3.2), one sees that at any point, the fluid is rotating
counterclockwise around the z-axis. Using the right-hand-rule, we thus expect the curl Ò ◊ v
to point in the positive z-direction, since it is the axis of rotation. We calculate:
Ò ◊ v = (0, 0, 2) .
Indeed, it points in the positive z-direction, as expected. Furthermore, half the magnitude of
the curl is
1
|Ò ◊ v| = 1,
2
which means that the angular speed of rotation of the ball would be 1 radian per unit of time.
⇤
Example 4.5.5 An irrotational velocity field. Conside a moving fluid with velocity field
given by
v(x, y, z) = (x, y, z).
Sketching the vector field, we see that this would be a fluid in expansion. As the fluid is
expanding, regardless of where the small sphere is located, it should not cause it to rotate (try
to visualize this yourself). So we expect the velocity field to be curl-free. From the definition
of the curl, we calculate:
Ò ◊ v = (0, 0, 0).
Thus it is curl-free, as expected, and this is an example of an irrotational velocity field. ⇤
Example 4.5.6 Another irrotational velocity field. Irrotational fluids do not have to
be necessarily spherically symmetric. Consider for instance a moving fluid with velocity field
This would be a fluid that is moving uniformly in the positive z-direction. If you think about
it for a little bit, it should not induce any rotation either, regardless of where the small sphere
is located. So we expect the velocity field to be curl-free again. Indeed, from the definition we
Ò ◊ v = (0, 0, 0), which says that the velocity field is irrotational. ⇤
As we have seen (see Exercise 2.1.3.2), the fluid is rotating counterclockwise around the z-axis.
It is not so obvious to see whether more fluid is existing or entering small spheres in the fluid.
It is easiest to consider first a sphere centered around the origin. Because of the rotational
motion, we see that actually no fluid is entering or leaving the sphere at all. So we expect the
divergence to be zero, at least at the origin.
It is not so obvious to see why the same should be true for all spheres not centered at the
origin, but you can try to visualize it. In the end, through direct calculation, we get that
Ò · v = 0, and hence the velocity field is divergence-free (that is, incompressible). ⇤
Example 4.5.9 Another incompressible velocity field. Consider the fluid with velocity
field
v(x, y, z) = (0, 0, 1),
which has uniform velocity in the positive z-direction. Here, if you pick a small sphere centered
anywhere, there is certainly fluid entering and exiting the sphere because of the linear motion
CHAPTER 4. k-FORMS 104
of the fluid, but the exact same amount of fluid will enter and exit the sphere. Therefore we
expect the divergence to be zero everywhere, and indeed, Ò · v = 0. Another example of an
incompressible velocity field.
Comparing with Example 4.5.6, we see that this velocity field is both irrotational and
incompressible, since it is both curl-free and divergence-free. ⇤
4.5.4 Exercises
1. Show that any vector field of the form
is irrotational.
Solution. To show that F is irrotational, we need to show that
Ò ◊ F = 0.
We can use directly the formula for the curl, or we can use the language of differential
forms. In the latter, we associate a one-form Ê to the vector field F:
We want to show that dÊ = 0. But this is clear true, as taking the exterior derivative
gives
df dg dh
dÊ = dx · dx + dy · dy + dz · dz = 0,
dx dy dz
since f (x), g(y), h(z) are only functions of x, y, z respectively. Therefore F is irrotational.
2. Show that any vector field of the form
is incompressible.
Solution. To show that F is incompressible, we need to show that
Ò · F = 0.
(a) Is Ò · F positive, negative, or zero at the origin? What about at the point (5, 0, 0)?
And (≠5, 0, 0)?
CHAPTER 4. k-FORMS 105
(a) Is Ò · F positive, negative, or zero at the origin? What about at (0, 5, 0)?
5. Consider a vector field F(x, y, z) = (f1 (x, y), f2 (x, y), 0) on R3 ; it is independendent of z,
and its z-component is zero. A sketch of the vector field in the xy-plane is shown in the
figure below; as it is independent of z, it looks the same in all other horizontal planes.
(a) Is Ò · F positive, negative, or zero at the origin? What about at (0, 5, 0)?
6. Consider a vector field F(x, y, z) = (f1 (x, y), f2 (x, y), 0) on R3 ; it is independendent of z,
and its z-component is zero. A sketch of the vector field in the xy-plane is shown in the
figure below; as it is independent of z, it looks the same in all other horizontal planes.
(a) Is Ò · F positive, negative, or zero at the origin? What about at (0, 5, 0)?
Objectives
You should be able to:
• Determine when a k-form is closed or exact, focusing on one-, two-, and three-forms in
R3 .
• Show that the general definition of closeness for one-forms reduces to our previous
definition in terms of partial derivatives.
which is precisely the condition stated in Definition 2.2.14. In fact, this explains where this
strange condition comes from! ⇤
As for one-forms, it is easy to show that exact forms are always closed: it follows directly
from the fact that d2 = 0, which is a key property of the exterior derivative proved in
Lemma 4.3.9.
Lemma 4.6.3 Exact k-forms are closed. If a k-form Ê on U ™ Rn is exact, then it is
closed.
Proof. This is a direct consequence of the fact that d2 = 0. If Ê is exact, then there exists a
(k ≠ 1)-form ÷ such that Ê = d÷. But then
dÊ = d(d÷) = 0
which is defined on U = R3 \ {(0, 0, 0)}, which is simply connected. One can check that Ê is
closed, but not exact. Does that contradict Poincare’s lemma? Fortunately it doesn’t, as U
is not an open ball in R3 (as it does not contain the origin). But it shows that there exists
k-forms with k Ø 2 defined on simply connected open sets that are closed but not exact.
CHAPTER 4. k-FORMS 111
Remark 4.6.7 The world of cohomology (this is just for fun and beyond the
scope of this class!). In fact, the relation between closed and exact forms is quite deep.
As we have seen, it is closely connected to the existence of “holes” in a space, which is the
subject of topology. In fact, studying when closed forms are not exact gives rise to the topic of
cohomology, which is an important branch of geometry and topology. Believe me, cohomology
is all over the place. You wouldn’t believe it, but it is even used to describe gauge theories in
physics!
While describing cohomology is way beyond the scope of this course, let me explain in a few
words what it is about, just for fun, in the context of differential forms. Suppose that Ê and ÷
are closed k-forms on U . If they differ from each other by an exact form, that is Ê ≠ ÷ = dfl
for some (k ≠ 1)-form fl, then we say that Ê and ÷ are “equivalent” (or “cohomologous”).
In this way, we construct equivalence classes of closed k-forms that differ by an exact form.
Those equivalence classes (which are called “cohomology classes”) form a vector space, which
is called the “de Rham cohomology space” H k (U ).1
If U is the whole of Rn , or an open ball in Rn , then all closed k-forms are exact, and hence
they are all equivalent to the zero k-form. The de Rham cohomology spaces H k (U ) (with
k Ø 1) are then all trivial (the zero vector space). So the de Rham cohomology spaces H k (U )
with k Ø 1 control, in a sense, how topologically non-trivial U is.
For instance, if we consider U = R2 \ {(0, 0)}, one can show that
H 1 (U ) = R, H 2 (U ) = 0.
The fact that H 2 (U ) = 0 says that all (closed) two-forms on U are exact. However, the
interesting fact here is that H 1 (U ) = R, which says that not all closed one-forms are exact;
indeed, we saw one example of such a one-form in Example 2.2.13, which was closed but not
exact. Since H 1 (U ) is one-dimensional, this means that all closed one-forms that are not
exact differ from the one we saw in that example by an exact form (they are in the same
cohomology class), up to overall rescaling.
F = Ò ◊ G.
3. The fact that exact two-forms are closed is the statement that if there exists a vector
potential for F, then F is divergence-free, namely Ò · F = 0. We call this the “screening
test for vector potentials”. However, while vector fields with a vector potential are
divergence-free, divergence-free vector fields do not necessarily have a vector potential.
4. Poincare’s lemma translates into the statement that if F is defined and has continuous
first order partial derivatives on all of R3 (or an open ball therein), then F has a vector
potential if and only if it is divergence-free.
Remark 4.6.8 Suppose that a vector field F on R3 has a vector potential G. This means
that there exists a vector field G such that
F = Ò ◊ G.
Finding G is not always obvious, as one would need in principle to integrate fairly complicated
partial differential equations. Moreover, G is far from unique.
It turns out that there is a nice result that drastically simplifies calculations. One can
show that, if F has a vector potential G, then we can always choose G to have vanishing
z-component function. In other words, if F has a vector potential, then there always exists a
G = (g1 , g2 , 0) such that
F = Ò ◊ G.
This is very helpful when trying to find a vector potential.
In the language of differential forms, this corresponds to the statement that if Ê is an
exact two-form on R3 , then there exists a one-form ÷ = f dx + g dy (with vanishing third
CHAPTER 4. k-FORMS 113
component function) such that Ê = d÷. We prove this statement in Exercise 4.6.3.6.
4.6.3 Exercises
CHAPTER 4. k-FORMS 114
Ê = xy dy · dz + yz dz · dx + zx dx · dy.
CHAPTER 4. k-FORMS 115
Determine whether F has a vector potential G. If it does, find such a vector potential.
Solution. Since the component functions of F are smooth on all of R3 , Poincare’s
lemma applies. So F will have a vector potential if and only if it is divergence-free. We
calculate its divergence:
ˆ ˆ ˆ
Ò·F= (yz) + (xz) + (xy)
ˆx ˆy ˆz
=0,
and thus we conclude that there exists a vector potential G such that Ò ◊ G = F.
Now we need to find G. We assume that G = (g1 , g2 , 0). Then the condition that
Ò ◊ G = F is
3 4
ˆg2 ˆg1 ˆg2 ˆg1
Ò◊G= ≠ , , ≠ = (yz, xz, xy).
ˆz ˆz ˆx ˆy
So we get three equations to solve. We integrate the first one (equality of the x-component
functions) to get:
yz 2
g2 (x, y, z) = ≠ + –(x, y).
2
We integrate the second one (equality of the y-component functions) to get:
xz 2
g1 (x, y, z) = + —(x, y).
2
Substituting in the third one (equality of the z-component functions), we get:
ˆg2 ˆg1 ˆ–(x, y) ˆ—(x, y)
≠ = ≠ = xy.
ˆx ˆy ˆx ˆy
We need to find any two functions – and — that satisfy this condition. We choose — = 0,
2
– = x2y . As a result, we have found a vector potential
A B
xz 2 y 2
G(x, y, z) = , (x ≠ z 2 ), 0 .
2 2
CHAPTER 4. k-FORMS 116
Similarly,
ˆ y (x2 + y 2 + z 2 ) ≠ 3y 2
=
ˆy (x2 + y 2 + z 2 )3/2 (x2 + y 2 + z 2 )5/2
and
ˆ z (x2 + y 2 + z 2 ) ≠ 3z 2
= .
ˆz (x2 + y 2 + z 2 )3/2 (x2 + y 2 + z 2 )5/2
Now, we get:
ˆ x ˆ y
dÊ = dx · dy · dz + dy · dz · dx
ˆx (x + y + z )
2 2 2 3/2 ˆy (x + y + z 2 )3/2
2 2
ˆ z
+ dz · dx · dy
ˆz (x2 + y 2 + z 2 )3/2
3(x2 + y 2 + z 2 ) ≠ 3x2 ≠ 3y 2 ≠ 3z 2
= dx · dy · dz
(x2 + y 2 + z 2 )5/2
=0.
(d) Does F have a vector potential? Justify your answer. If it does, find such a vector
potential.
CHAPTER 4. k-FORMS 117
Solution.
(a) F is defined (and in fact, is smooth) wherever the denominator is non-zero. This is
wherever (x, y) ”= (0, 0). So the domain of definition of F is
This is R3 minus the z-axis. It is path connected, since any two points in U can
be connected by a path. It is however not simply connected, since a closed curve
around the z-axis cannot be continuously contracted to a point within U (it would
hit the z-axis, which is not in U ).
So F is curl-free.
(d) We found in (b) that the divergence of F is non-zero. This means that F cannot have
a vector potential, since vector fields that have a vector potential are divergence-free.
(e) We found in (c) that F is curl-free, so it passes the screening test for conservative
vector fields. However, since its domain of definition U is not simply connected,
Poincare’s lemma does not apply. We thus cannot conclude whether F is conserva-
tive from the statement that it is curl-free.
In fact, we can show that it is not conservative by showing that its integral along a
closed loop is non-zero, as in Exercise 3.4.3.2. Let Ê be the one-form associated to
F:
y x
Ê=≠ 2 dx + 2 dy + z dz.
x + y2 x + y2
CHAPTER 4. k-FORMS 118
Consider the parametric curve – : [0, 2fi] æ R3 with –(t) = (cos(t), sin(t), 0), which
is the unit circle (counterclockwise) around the origin in the xy-plane. The pullback
of Ê is:
3 4
sin(t) cos(t)
–ú Ê = ≠ 2 (≠ sin(t)) + cos(t) dt
cos (t) + sin (t)
2 cos (t) + sin2 (t)
2
=dt.
Since this is non-zero, by Corollary 3.4.3 (or, in other words, the Fundamental
Theorem for line integrals), Ê cannot be exact, since the line integrals of exact
one-forms along closed curves vanish. Equivalently, the vector field F is not
conservative.
6. Let Ê be an exact two-form on R3 . Show that there exists a one-form ÷ of the form
÷ = f dx + g dy
(i.e. with a vanishing z-component function) such that d÷ = Ê.
In other words, in the language of vector calculus, if a vector field F has a vector
potential G, then G can always be chosen to take the form G = (g1 , g2 , 0).
Solution. Since Ê is exact, we know that there exists a one-form — such that d— = Ê.
Suppose that we find such a —:
— = b1 dx + b2 dy + b3 dz,
for some component functions b1 , b2 , b3 . — is certainly not unique; there are many one-
forms such that their exterior derivative equals Ê. In fact, since d2 = 0, we can add to —
any exact one-form dF for a function F , and we get another one-form whose exterior
derivative is Ê. That is, if we define
—˜ = — ≠ dF
for any function F , then d—˜ = d— ≠ d2 F = Ê.
In particular, since
ˆF ˆF ˆF
dF = dx + dy + dz,
ˆx ˆy ˆz
if we can choose F such that
ˆF
= b3 ,
ˆz
then we see that 3 4 3 4
˜ ˆF ˆF
— = b1 ≠ dx + b2 ≠ dy,
ˆx ˆy
and hence it is of the form that we are looking for (no z-component function). But this
is easy to do; simply pick
⁄
F (x, y, z) = b3 (x, y, z) dz,
CHAPTER 4. k-FORMS 119
i.e. any antiderivative in the z-variable will do. So we conclude that we can always find
a one-form —˜ with no z-component and such that d—˜ = Ê.
Objectives
You should be able to:
• Show that the pullback of a 2-form in R2 and a 3-form in R3 is given by the Jacobian
determinant.
2. The pullback of a product of a zero-form and a one-form is the product of the pullbacks.
3. The pullback of the exterior derivative of a zero-form is the exterior derivative of the
pullback.
In this section we define the pullback of a k-form using a similar approach. All we need to do
is generalize the first and second properties to k-forms.
More precisely, let „ : V æ U be a smooth function, with V ™ Rm and U ™ Rn open
subsets. We want the pullback „ú to satisfy the following properties:
1. If Ê and ÷ are k-forms on U , then
„ú (Ê + ÷) = „ú Ê + „ú ÷.
„ú (df ) = d(„ú f ).
CHAPTER 4. k-FORMS 120
The third property is unchanged, and the first two are naturally generalized to k-forms.
Imposing these properties gives us a unique definition for the pullback of a k-form.
Lemma 4.7.1 The pullback of a k-form. Let „ : V æ U be a smooth function, with V ™
Rm and U ™ Rn open subsets. We write t = (t1 , . . . , tm ) œ V , and „(t) = (x1 (t), . . . , xn (t)).
Let Ê be a k-form on U , with
ÿ
Ê= fi1 ···ik dxi1 · · · · · dxik ,
1Æi1 <...<ik Æn
for some functions fi1 ···ik : U æ R. Then the pullback „ú Ê is a k-form on V given by:
ÿ 3 4
ˆxi1 ˆxi1
„ Ê=
ú
fi1 ···ik („(t)) dt1 + . . . + dtm
1Æi1 <...<ik Æn
ˆt1 ˆtm
3 4
ˆxik ˆxik
· ··· · dt1 + . . . + dtm .
ˆt1 ˆtm
Proof. Start with a basic k-form
dxi1 · · · · · dxik .
Since we impose (Property 2) that „ú (Ê · ÷) = („ú Ê) · „ú ÷), we know that
But we already calculated how basic one-forms transform in Lemma 2.4.4; they transform as
differentials would. This calculation followed from Property 3, which we still impose here, so
it is still valid. This gives us the pullback of basic k-forms.
Then we use Property 2 to extended it to a function times a basic k-form, and then
Property 1 to extend it to linear combination of such terms, to get the final result for the
pullback of a generic k-form. ⌅
This formula certainly looks ugly, but it is really not that bad. Concretely, all you need to
do is compose the component functions with „, and transform the basic one-forms one by one
in the wedge product as in Lemma 2.4.4, i.e. they transform as you would expect differentials
transform. That’s it! This will certainly be clearer with examples.
Example 4.7.2 The pullback of a two-form. Consider the following two-form on R3 :
Ê = xy dy · dz + (xz + y) dz · dx + dx · dy,
and the smooth function „ : R2 æ R3 given by
„(u, v) = (uv, u2 , v 2 ) = (x(u, v), y(u, v), z(u, v)).
To calculate „ú Ê, let us start by calculating the pullback of the basic one-forms. We get:
ˆx ˆx
„ú (dx) = du + dv = v du + u dv,
ˆu ˆv
ˆy ˆy
„ú (dy) = du + dv = 2u du,
ˆu ˆv
ˆz ˆz
„ú (dz) = du + dv = 2v dv.
ˆu ˆv
CHAPTER 4. k-FORMS 121
„ú Ê =(uv)(u2 )„ú (dy) · „ú (dz) + ((uv)(v 2 ) + u2 )„ú (dz) · „ú (dx) + „ú (dx) · „ú (dy)
=u3 v(2u du) · (2v dv) + (uv 3 + u2 )(2v dv) · (v du + u dv) + (v du + u dv) · (2u du)
=4u4 v 2 du · dv + 2uv 2 (v 3 + u) dv · du + 2u2 dv · du
=(4u4 v 2 ≠ 2uv 5 ≠ 2u2 v 2 ≠ 2u2 )du · dv.
Ê = ex+y+z dx · dy · dz,
where we used the fact that the basic three-forms vanish whenever one of the factor is repeated,
and dv · dw · du = du · dv · dw since we need to exchange two basic one-forms twice to
relate the two basic three-forms. ⇤
Good, so we are now experts at computing pullbacks! Calculating the pullback of a k-form
is not more difficult than calculating the pullback of a one-form, but the calculation may be
longer, and you need to use the anti-symmetry properties of basic k-forms in Lemma 4.1.6 to
simplify the result at the end of your calculation.
Property 3 in our axiomatic definition states that the pullback commutes with the exterior
derivative for zero-form. It turns out that this property, which is very important, holds in
general for k-forms. Let us prove that.
Lemma 4.7.4 The pullback commutes with the exterior derivative. Let „ : V æ U
be a smooth function, with V ™ Rm and U ™ Rn open subsets. Let Ê be a k-form on U . Then:
since the fi1 ···ik are just functions (zero-forms). Thus we have:
ÿ
„ú (dÊ) = d(„ú fi1 ···ik ) · „ú (dxi1 ) · · · · · „ú (dxik )
1Æi1 <···<ik Æn
Q R
ÿ
=d a („ú fi1 ···ik )„ú (dxi1 ) · · · · · „ú (dxik )b
1Æi1 <···<ik Æn
=d(„ Ê),ú
Definition 4.7.6 Top form. We call an n-form on an open subset U ™ Rn a top form. ⌃
Such forms are called “top forms” because all forms with k Ø n necessarily vanish on Rn .
Going back to the pullback, we get the nice following result when we pullback a top form:
Lemma 4.7.7 The pullback of a top form in Rn in terms of the Jacobian determinant.
Suppose that Ê is a top form on U ™ Rn , and „ : V æ U a smooth function with V ™ Rn . As
above we write
„(t) = (x1 (t), . . . , xn (t))
with t = (t1 , . . . , tn ) œ V , and
Ê = f dx1 · . . . · dxn
for some function f : U æ R. Then
Proof for R2 and R3 . It is not so easy to write a general proof for Rn using the computational
approach for the pullback that we have used so far. We will be able to write down a general
proof easily in the next subsection after having introduced a more algebraic approach to the
pullback. For the time being, let us prove the statement explicitly for R2 and R3 .
For R2 , our function „ takes the form
where A B
ˆx1 ˆx1
ˆx1 ˆx2 ˆx2 ˆx1
det J„ = det ˆt1
ˆx2
ˆt2
ˆx2 = ≠ .
ˆt1 ˆt2 ˆt1 ˆt2 ˆt1 ˆt2
But
3 4 3 4
ˆx1 ˆx1 ˆx2 ˆx2
„ú (dx1 · dx2 ) = dt1 + dt2 · dt1 + dt2
ˆt1 ˆt2 ˆt1 ˆt2
3 4
ˆx1 ˆx2 ˆx2 ˆx1
= ≠ dt1 · dt2 ,
ˆt1 ˆt2 ˆt1 ˆt2
and the lemma is proved.
The calculation is similar but more involved for R3 . We have:
where
Q ˆx1 ˆx1 ˆx1 R
ˆt ˆt2 ˆt3
c 12 ˆx2 d
det J„ = det a ˆx
ˆt1
ˆx2
ˆt2 ˆt3 b
ˆx3 ˆx3 ˆx3
ˆt1 ˆt2 ˆt3
ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3
= + +
ˆt1 ˆt2 ˆt3 ˆt2 ˆt3 ˆt1 ˆt3 ˆt1 ˆt2
ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3
≠ ≠ ≠ .
ˆt2 ˆt1 ˆt3 ˆt1 ˆt3 ˆt2 ˆt3 ˆt2 ˆt1
But
3 4
ˆx1 ˆx1 ˆx1
„ú (dx1 · dx2 · dx3 ) = dt1 + dt2 + dt3
ˆt1 ˆt2 ˆt3
3 4
ˆx2 ˆx2 ˆx2
· dt1 + dt2 + dt3
ˆt ˆt2 ˆt3
3 1 4
ˆx3 ˆx3 ˆx3
· dt1 + dt2 + dt3
ˆt1 ˆt2 ˆt3
1 ˆx ˆx ˆx ˆx ˆx ˆx ˆx1 ˆx2 ˆx3
1 2 3 1 2 3
= + +
ˆt1 ˆt2 ˆt3 ˆt2 ˆt3 ˆt1 ˆt3 ˆt1 ˆt2
ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3 ˆx1 ˆx2 ˆx3 2
≠ ≠ ≠ dt1 · dt2 · dt3 ,
ˆt2 ˆt1 ˆt3 ˆt1 ˆt3 ˆt2 ˆt3 ˆt2 ˆt1
is invertible).
This is a very powerful theorem, which relates invertibility of a function to invertibility of
its Jacobian matrix.
Another important theorem that involves the Jacobian is the Implicit Function Theorem.
This theorem concerns the following question: suppose that we are given a number of relations
between a set of variables, can we represent these relations as the graph of a function? Let us
be more precise.
Start with single variable calculus. Let f : R2 æ R be a C 1 -function, which we write as
f (x, y). Setting f (x, y) = 0 gives a relation between the variables x and y. Can we think
of this relation as implicitly defining y as a function of x? The answer is yes, under mild
conditions, and only locally.
More precisely, the statement is that, for any point (a, b) in the domain of f such that
f (a, b) = 0, if ˆf
ˆy (a, b) ”= 0, then there exists an open neighborhood U ™ R of a such that
there exists a unique C 1 -function g : U æ R such that g(a) = b and f (x, g(x)) = 0 for all
x œ U . In other words, if ˆf ˆy (a, b) ”= 0, the relation f (x, y) = 0 implicitly and uniquely defines
1
y as a C -function y = g(x) locally near the point (a, b).
This statement can be naturally generalized to multivariable calculus using the Jacobian
determinant.
Theorem 4.7.9 Implicit Function Theorem. Let f : Rn+m æ Rm be a C 1 -function. Let
us denote by (x, y) = (x1 , . . . , xn , y1 , . . . , ym ) coordinates on Rn+m . We can write the function
f in terms of its component functions as
Let (a, b) œ Rn+m be a point such that f (a, b) = 0. Suppose that det Jf,y (a, b) ”= 0, where
Jf,y is the Jacobian of f with respect to the variables y:
Q ˆf1 ˆf1 R
ˆy1 ··· ˆym
c . .. .. d
a ..
Jf,y = c . d
. b.
ˆfm ˆfm
ˆy1 ··· ˆym
(In other words, this Jacobian matrix is invertible.) Then there exists an open neighborhood
U ™ Rn of a such that there exists a unique C 1 -function g : U æ Rm such that g(a) = b and
f (x, g(x)) = 0 for all x œ U .
This is the natural multivariable generalization. If the Jacobian matrix of the function
f (x, y) with respect to the y-variables is invertible at a point (a, b), the relation f (x, y) = 0
implicitly and uniquely defines the variables y as C 1 -functions y = (g1 (x), . . . , gm (x)) locally
near (a, b).
Recall from Definition 4.1.1 (naturally generalized to Rn ) that a basic one-form dxi is a
linear map dxi : Rn æ R which acts as
dxi (u1 , . . . , un ) = ui ,
i.e. it outputs the i’th component of the vector u œ Rn . As we are now thinking of the basic
one-forms dxi as linear maps, we can define their pullback by composition, as we originally
did for functions in Definition 2.4.1.
We first define and study the pullback when „ is a linear map between vector spaces. We
will generalize to the case of a smooth function afterwards.
Definition 4.7.10 The pullback of a basic one-form with respect to a linear map.
Let dxi : Rn æ R be a basic one-form, and let „ : Rm æ Rn be a linear map. We define the
pullback „ú (dxi ) : Rm æ R by composition:
„ú (dxi ) = dxi ¶ „ : Rm æ R.
⌃
Let us now give an explicit formula for the pullback of a basic one-form.
Lemma 4.7.11 An explicit formula for the pullback of a basic one-form. Let
dxi : Rn æ R, i = 1, . . . , n be the basic one-forms on Rn , and dtj : Rm æ R, j = 1, . . . , m
be the basic one-forms on Rm . The linear map „ : Rm æ Rn can be represented by a matrix
A = (aij ). Then we can write
m
ÿ
„ú (dxi ) = aij dtj = ai1 dt1 + . . . + aim dtm .
j=1
⌅
It is then easy to generalize the definition of pullback to the basic k-forms.
Definition 4.7.12 The pullback of a basic k-form with respect to a linear map.
Let dxi1 · . . . · dxik : (Rn )k æ R be a basic k-form, and let „ : Rm æ Rn be a linear map.
Let v1 , . . . , vk œ Rm be vectors. We define the pullback „ú (dxi1 · . . . · dxik ) : (Rm )k æ R by:
⌃
It follows directly from the definition that the pullback commutes with the wedge product:
CHAPTER 4. k-FORMS 127
Lemma 4.7.13 The pullback commutes with the wedge product. Under the setup
above,
„ú (dxi1 · . . . · dxik ) = „ú (dxi1 ) · . . . · „ú (dxik ).
Proof. By definition,
⌅
As a corollary, we obtain an explicit formula to calculate the pullback of a basic k-form.
Corollary 4.7.14 An explicit formula for the pullback of a basic k-form. Let
dxi1 · . . . · dxik : (Rn )k æ R be a basic k-form, and let „ : Rm æ Rn be a linear map. Let
dtj : Rm æ R, j = 1, . . . , m be the basic one-forms on Rm . The linear map „ : Rm æ Rn can
be represented by a matrix A = (aij ). Then we can write
„ú (dxi1 · . . . · dxik ) = (ai1 1 dt1 + . . . + ai1 m dtm ) · · · · · (aik 1 dt1 + . . . + aik m dtm ) .
We can also write down an explicit formula in terms of the determinant when we pullback
a basic n-form from Rn to Rn .
Lemma 4.7.15 The pullback of a basic n-form in Rn . Let dx1 · · · · · dxn : (Rn )n æ R,
and let „ : Rn æ Rn be a linear map, which can be representated by an n ◊ n matrix A = (aij ).
Then
„ú (dx1 · · · · · dxn ) = (det A) dx1 · · · · · dxn .
functions „ : V æ U with V ™ Rm ? The idea is simple. For any point t œ V , the Jacobian
matrix of „ (i.e. the matrix of first partial derivatives), also called the “total derivative of „”,
CHAPTER 4. k-FORMS 128
Then we see that we recover all the formulae that we obtained previously, and that our
three fundamental properties are satisfied! The pullback of a basic one-form becomes
ˆxi ˆxi
„ú (dxi ) = dt1 + . . . + dtn ,
ˆt1 ˆtn
as before. Property 1 is obviously satisfied by definition. Lemma 4.7.13 becomes Property 2,
and Corollary 4.7.14 becomes our general formula for the pullback of k-forms in Lemma 4.7.1.
It is easy to check that Property 3 is satisfied. Finally, we obtain a general proof of Lemma 4.7.7
on Rn , as this is simply Lemma 4.7.15. Neat!
4.7.5 Exercises
1. Consider the basic two-form dx · dy on R2 , and the polar coordinate transformation
– : R2 æ R2 with
–(r, ◊) = (r cos ◊, r sin ◊).
Show by explicit calculation that
Therefore,
–ú (dx · dy) = r dr · d◊ = (det J– )dr · d◊
as claimed.
2. Let
Ê = (x2 + y 2 + z 2 ) dx · dy · dz
on R3 , and – : R3 æ R3 the spherical transformation
Solution.
(a) We calculate the pullback:
Next we show that this det J– = r2 sin(◊). By definition of the Jacobian matrix,
we get:
Q R
ˆx ˆx ˆx
ˆr
c ˆy ˆ◊ ˆ„ d
det J– = det c
a ˆr
ˆy
ˆ◊
ˆy d
ˆ„ b
ˆz ˆz ˆz
ˆr ˆ◊ ˆ„
Q R
sin(◊) cos(„) r cos(◊) cos(„) ≠r sin(◊) sin(„)
c d
= det a sin(◊) sin(„) r cos(◊) sin(„) r sin(◊) cos(„) b
cos(◊) ≠r sin(◊) 0
=r2 sin(◊) cos2 (◊) cos2 („) + r2 sin3 (◊) sin2 („) + r2 sin3 (◊) cos2 („)
+ r2 sin(◊) cos2 (◊) sin2 („)
=r2 sin(◊).
Therefore
as claimed.
CHAPTER 4. k-FORMS 130
–ú Ê =(r2 sin2 (◊) cos2 („) + r2 sin2 (◊) sin2 („) + r2 cos2 (◊))–ú (dx · dy · dz)
=(r2 sin2 (◊) + r2 cos2 (◊))r2 sin(◊)dr · d◊ · d„
=r4 sin(◊)dr · d◊ · d„.
3. Let
Ê = (x2 + y 2 ) (x dy · dz + y dz · dx + z dx · dy)
on R3 , and – : R3 æ R3 be the cylindrical transformation
(a) Find –ú Ê.
Solution.
(a) To simplify the calculation of the pullback, let us first calculate the pullback of the
basic two-forms:
and
On the other hand, we can calculate first the exterior derivative of Ê. We get:
We conclude that
d(–ú Ê) = –ú (dÊ),
as claimed.
4. Let
Ê = zexy dx · dy
on R3 , and „ : (R”=0 )2 æ R3 with:
3 4
u v
„(u, v) = , , uv .
v u
Find „ú Ê.
Solution. We calculate the pullback:
3 4 3 4
1 u v 1
„ Ê =uve du ≠ 2 dv · ≠ 2 du + dv
u v
ú v u
v v u u
3 4
1 1
=uve ≠ du · dv
uv uv
=0.
5. Show that the pullback of an exact form is always exact.
Solution. Let Ê be an exact k-form on U ™ Rn , and let „ : V æ U be a smooth
function for V ™ Rm . We want to prove that the pullback of Ê is exact.
Since Ê is exact, we know that there exists a (k ≠ 1)-form ÷ on U such that Ê = d÷.
CHAPTER 4. k-FORMS 132
Then
„ú Ê = „ú (d÷) = d(„ú ÷),
where we used the fact that the pullback commutes with the exterior derivative (see
Lemma 4.7.4). Therefore, the k-form „ú Ê on V is the exterior derivative of a (k ≠ 1)-form
„ú ÷ on V , and hence it is exact.
6. Let Ê be a k-form on Rn , and Id : Rn æ Rn be the identity function defined by
Id(x1 , . . . , xn ) = (x1 , . . . , xn ). Show that
Idú Ê = Ê.
for some smooth functions fi1 ···ik : Rn æ R. We know that Idú (dxi ) = dxi for all
i œ {1, . . . , n}, by definition of the identity map. Similarly, for any function f : Rn æ R,
Idú f = f . As a result, we get:
ÿ
Idú Ê = Idú (fi1 ···ik )Idú (dxi1 ) · · · · · Idú (dxik )
1Æi1 <···<ik Æn
ÿ
= fi1 ···ik dxi1 · · · · · dxik
1Æi1 <···<ik Æn
=Ê.
7. Let Ê be a k-form on U ™ Rn . Let „ : V æ U and – : W æ V be smooth functions,
with V ™ Rm and W ™ R¸ . Show that
In other words, it doesn’t matter whether we pullback in one or two steps in the chain of
maps
– „
W æ V æ U.
for some smooth functions fi1 ···ik : U æ R. On the one hand, the pullback by „ ¶ – is
ÿ
(„ ¶ –)ú Ê = („ ¶ –)ú (fi1 ···ik )(„ ¶ –)ú (dxi1 ) · · · · · („ ¶ –)ú (dxik ).
1Æi1 <···<ik Æn
(„ ¶ –)ú f = –ú („ú f )
(„ ¶ –)ú f = f ¶ „ ¶ –,
while
–ú („ú f ) = –ú (f ¶ „) = f ¶ „ ¶ –.
Thus
(„ ¶ –)ú f = –ú („ú f ).
As for the basic one-forms, let us introduce further notation for the maps „ : V æ U
and – : W æ V . Let us write z = (z1 , . . . , z¸ ) for coordinates on W ; y = (y1 , . . . , ym )
for coordinates on V ; and x = (x1 , . . . , xn ) for coordinates on U . We write „(y) =
(„1 (y, . . . , „n (y)), and –(z) = (–1 (z), . . . , –m (z)). Then, we have:
m
ÿ ˆ„i
„ú dxi = dya ,
a=1
ˆya
and
m
ÿ̧ ÿ ˆ„i -- ˆ–a
–ú („ú dxi ) = - dzb .
b=1 a=1
ˆya y=–(z) ˆzb
On the other hand, if we pullback by the composition of the maps, we get:
ÿ̧ ˆ(„ ¶ –)i
(„ ¶ –)ú dxi = dzb .
b=1
ˆzb
Objectives
You should be able to:
Ê · ıÊ = dx1 · · · · · dxn .
To define the Hodge star dual of a general k-form on U ™ Rn , we apply the Hodge star to
each summand. More precisely, if
ÿ
÷= fi1 ···ik dxi1 · · · · · dxik
1Æi1 <...<ik Æn
⌃
To make sense of this definition, let us look at the Hodge star action on the basic k-forms
for low-dimensional space.
Example 4.8.2 The action of the Hodge star in R. There are only two basic k-forms
in R, namely the zero-form 1 and the one-form dx. From the definition, we want 1 · ı1 = dx
and dx · ıdx = dx, from which we conclude that:
ı1 = dx, ıdx = 1.
1
This standard notation should not be confused with the pullback of a differential form; those are very
different things.
CHAPTER 4. k-FORMS 135
The Hodge star thus provides a duality between zero-forms and one-forms in R. ⇤
Example 4.8.3 The action of the Hodge star in R2 . It becomes a little more interesting
in R2 . The basic forms are the zero-form 1, the one-forms dx and dy, and the two-form dx · dy.
From the definition, we want 1 · ı1 = dx · dy, dx · ıdx = dx · dy, dy · ıdy = dx · dy, and
(dx · dy) · ı(dx · dy) = dx · dy. We conclude that
It thus provides a duality between zero-forms and two-forms in R2 , and a “self-duality” for
one-forms. Note that the sign is important here for the action on the basic one-forms. ⇤
Example 4.8.4 The action of the Hodge star in R3 . Things become even more
interesting in R3 . The basic forms are the zero-form 1, the one-forms dx, dy, dz, the two-forms
dy · dz, dz · dx, dx · dy, and the three-form dx · dy · dz. From the definition, we get that:
Thus, in R3 , it provides a duality between zero-forms and three-forms, and between one-forms
and two-forms. ⇤
Example 4.8.5 An example of the Hodge star action in R3 . Consider the two-form
Ê = xyz dy · dz + ex dx · dy. Its Hodge star dual is the one-form:
⇤
The action of the Hodge star in R3 naturally justifies our dictionary to translate between
k-forms in R3 and vector calculus objects in Table 4.1.11. Indeed, let Ê = f dx + g dy + h dz
be a one-form on R3 . Then its Hodge dual is the two-form
ıÊ = f dy · dz + g dz · dx + h dx · dy.
This is why we used this particular choice for the basic two-forms in Table 4.1.11; it’s because
this is what one gets through Hodge duality, which identifies one-forms and two-forms in R3 .
In fact, what this means is that we really only needed the first two lines in Table 4.1.11.
Indeed, we can always transform a two-form into a one-form by taking its Hodge dual, and a
three-form into a zero-form. So, in the end, all that we need to establish a dictionary between
differential forms and vector calculus objects is to say that zero-forms are functions, and
one-forms correspond to vector fields.
For instance, if F is the vector field associated to a one-form Ê, we could have defined the
curl Ò ◊ F to be the vector field associated to the one-form ıdÊ. That is,
Ê¡F ı dÊ ¡ Ò ◊ F.
CHAPTER 4. k-FORMS 136
Similarly, we could have defined the divergence Ò · F to be the function given by ıd ı Ê. That
is,
Ê¡F ı d ı Ê ¡ Ò · F.
We could translate all vector calculus identities in Section 4.4 using the Hodge star, but in
the end, as far as we are concerned in this course, this is just a fancier way of saying the same
thing. :-)
Example 4.8.6 Maxwell’s equations using differential forms (optional). Recall
from Example 4.4.7 the statement of Maxwell’s equations, which form the foundations of
electromagnetism. They can be written in terms of the electric vector field E and the magnetic
vector field B on R3 as follows (I am now using units with c = 1 as is standard in modern
physics, and I have rescaled the electric charge fl and the electric current density J to absorb
the factor of 4fi):
Ò · E =fl,
Ò · B =0,
ˆB
Ò◊E+ =0,
ˆt
ˆE
Ò◊B≠ =J.
ˆt
It turns out that there is a very nice way of rewriting Maxwell’s equations using differential
forms, which makes them manifestly relativistic (i.e. consistent with special relativity).
Moreover, this reformulation works in any number of dimensions! It defines the natural
generalization of Maxwell’s equations to higher-dimensional spacetimes, which is useful in
physics theories like string theory.
To write Maxwell’s equations in this form, we need to consider them as living on spacetime,
i.e. R4 , with coordinates (t, x, y, z). We first construct a two-form F on R4 that combines the
electric field E and the magnetic field B as follows:
F = Bx dy · dz + By dz · dx + Bz dx · dy + Ex dx · dt + Ey dy · dt + Ez dz · dt.
We also construct a three-form which combines the electric current J and the electric charge
fl as:
J = fl dx · dy · dz ≠ jx dt · dy · dz ≠ jy dt · dz · dx ≠ jz dt · dx · dy.
Using these definitions, and the definition of the Hodge star on R4 ,2 we can rewrite Maxwell’s
equations neatly as the following two equations (you can check this!):
dF =0,
d ı F =J.
Isn’t that neat? The first equation is simply saying that the two-form F is closed. The second
equation is the “source equation”; if there is no source (i.e. J = 0), it is simply saying that
the two-form ıF is also closed. Moreover, not only is this formulation nice and clean, but it is
manifestly Lorentz invariant (as it is formulated in four-dimensional space time).
Furthermore, this formulation of Maxwell’s equation naturally generalizes to any number
of dimensions. In Rn , F remains a two-form, but J becomes an (n ≠ 1)-form. Then ıF
CHAPTER 4. k-FORMS 137
is an (n ≠ 2)-form, and the two equations make sense in Rn . As mentioned above, this
higher-dimensional generalization is useful in modern physical theories such as string theory.
This is an example of the power of the formalism of differential forms! ⇤
Ê = ≠(d” + ”d)Ê,
Ê · ıÊ = ±dt · dx · dy · dz,
with a plus sign whenever Ê does not contain dt, and a minus sign whenever Ê contains dt.
CHAPTER 4. k-FORMS 138
which is also a one-form. Let us now extract its associated vector field. Let F be the vector
field associated to Ê.
We look at the first term on the right-hand-side. ıÊ is a two-form associated to F. d ı Ê
then takes the divergence Ò · F. ıd ı Ê maps this to a zero-form, and then d ı d ı Ê take
the gradient of the resuling zero-form. The result is that the vector field associated to the
one-form d ı d ı Ê is
Ò(Ò · F).
Let us now look at the second term on the right-hand-side. dÊ takes the curl Ò ◊ F. ıdÊ
then maps it to a one-form associated to the vector field Ò ◊ F. d ı dÊ then takes the curl
again, Ò ◊ (Ò ◊ F), and finaly ıd ı dÊ maps it back to a one-form. The result is that the
vector field associated to the one-form ıd ı dÊ is
Ò ◊ (Ò ◊ F).
Ò(Ò · F) ≠ Ò ◊ (Ò ◊ F),
which we can take as the definition of the Laplacian of the vector field F. To show that it
takes the form
ˆ2F ˆ2F ˆ2F
Ò2 F = + + ,
ˆx2 ˆy 2 ˆz 2
one only needs to do an explicit calculation in R3 , see Exercise 4.8.4.5. ⌅
CHAPTER 4. k-FORMS 139
1.
Ò2 (f g) = f Ò2 g + 2Òf · Òg + gÒ2 f,
2.
Ò · (f Òg ≠ gÒf ) = f Ò2 g ≠ gÒ2 f.
Proof. These two identities follow from the graded product rule for the exterior derivative.
dF =0,
d ı F =J.
Ò2 (f g) = (f g)
= ı d ı d(f g)
= ı d ı (g df + f dg)
= ı d(g ı df + f ı dg)
= ı (dg · ıdf + g d ı df + df · ıdg + f d ı dg)
=Òg · Òf + gÒ2 f + Òf · Òg + f Ò2 g
=f Ò2 g + 2Òf · Òg + gÒ2 f.
In the proof we used the fact that Ò2 f = ıd ı df as in the proof of Lemma 4.8.8.
For the second identity, we get the following:
Ò · (f Òg ≠ gÒf ) = ı d ı (f dg ≠ gdf )
= ı d(f ı dg ≠ g ı df )
= ı (df · ıdg + f d ı dg ≠ dg · ıdf ≠ gd ı df )
To proceed we need to use a result which we haven’t proved. For any two k-forms Ê and
÷ on R3 , there’s a general result that says that
Ê · ı÷ = ıÊ · ÷.
Note that it is important that Ê and ÷ are both k-forms (same k), otherwise it wouldn’t apply.
It it not difficult to prove this statement, but since we do not need it anywhere else, we leave
the proof as an exercise (see Exercise 4.8.4.4).
Now in our previous expression we had the terms df · ıdg and ≠dg · ıdf . Since df and
dg are both one-forms,
df · ıdg = ıdf · dg,
CHAPTER 4. k-FORMS 140
4.8.4 Exercises
1. Let Ê be the two-form
Ê = xy dy · dz + xyz dz · dx + y dx · dy
on R3 . Find ıÊ.
Solution. We calculate the one-form ıÊ using the action of the Hodge star on basic
two-forms in R3 :
ı ı Ê = (≠1)k(n≠k) Ê.
Solution. We only need to prove the statement for basic k-forms as by definition of the
action of the Hodge star operator in Definition 4.8.1 it will then follow for all k-forms.
Let – be a basic k-form on Rn . By definition of the Hodge star, we know that ı– is
the unique basic (n ≠ k)-form such that
– · ı– = dx1 · · · · · dxn .
Now consider the basic (n ≠ k)-form ı–. By definition of the Hodge star, ı ı – will be
the unique basic k-form such that
ı– · ı ı – = dx1 · · · · · dxn .
– · ı– = ı– · ı ı –.
Using graded commutativity of the wedge product as in Lemma 4.2.6, we can rewrite
the right-hand-side as:
– · ı– = (≠1)k(n≠k) ı ı– · ı–.
Finally, given ı–, we know that the left-hand-side uniquely defines – (by definition of
the Hodge star), while the right-hand-side uniquely deefines ı ı – (again by definition of
the Hodge star), and therefore we must have
– = (≠1)k(n≠k) ı ı–.
CHAPTER 4. k-FORMS 141
3. Let Ê and ÷ be one-forms on Rn , and F and G be the associated vector fields. Show that
ı(Ê · ı÷) = F · G.
where we used the definition of the Hodge star for basic one-forms. Finally, since
ı(dx1 · · · · dxn ) = 1, we get:
n
ÿ n
ÿ
ı(Ê · ı÷) = fi gi ı (dx1 · · · · · dxn ) = fi gi = F · G,
i=1 i=1
where the last equality is for the associated vector fields F = (f1 , . . . , fn ) and G =
(g1 , . . . , gn ).
4. Let Ê and ÷ be k-forms on Rn . Show that
Ê · ı÷ = (≠1)k(n≠k) ı Ê · ÷.
Note that it is important that Ê and ÷ are both k-forms (same k), otherwise this property
wouldn’t apply.
Solution. The proof is similar in spirit to the solution of the previous problem. Since Ê
and ÷ are both k-forms on Rn , we can write both as linear combinations of basic k-forms
in Rn :
ÿ
Ê= fi1 ···ik dxi1 · · · · · dxik ,
1Æi1 <···<ik Æn
ÿ
÷= gi1 ···ik dxi1 · · · · · dxik ,
1Æi1 <···<ik Æn
Similarly, we have:
ÿ
ıÊ · ÷ = fi1 ···ik gi1 ···ik ı (dxi1 · · · · · dxik ) · dxi1 · · · · · dxik
1Æi1 <···<ik Æn
ÿ
=(≠1)k(n≠k) fi1 ···ik gi1 ···ik dxi1 · · · · · dxik · ı(dxi1 · · · · · dxik ),
1Æi1 <···<ik Æn
where we used graded commutativity of the wedge product, Lemma 4.2.6. We thus
conclude that
Ê · ı÷ = (≠1)k(n≠k) ı Ê · ÷.
5. Let F = (f1 , f2 , f3 ) be a smooth vector field in R3 . Show that
ˆ2F ˆ2F ˆ2F
Ò(Ò · F) ≠ Ò ◊ (Ò ◊ F) = + + .
ˆx2 ˆy 2 ˆz 2
This is the definition of the Laplacian of the vector field Ò2 F as in Lemma 4.8.9.
Solution. This is just an explicit and rather painful calculation. Let us do it step-by-
step. First,
ˆf1 ˆf2 ˆf3
Ò·F= + + .
ˆx ˆy ˆz
Thus
A B
ˆ 2 f1 ˆ 2 f2 ˆ 2 f3 ˆ 2 f1 ˆ 2 f2 ˆ 2 f3 ˆ 2 f1 ˆ 2 f2 ˆ 2 f3
Ò(Ò · F) = + + , + + , + + .
ˆx2 ˆxˆy ˆxˆz ˆyˆx ˆy 2 ˆyˆz ˆzˆx ˆzˆy ˆz 2
Next, we move on to the curl. First, we have
3 4
ˆf3 ˆf2 ˆf1 ˆf3 ˆf2 ˆf1
Ò◊F= ≠ , ≠ , ≠ .
ˆy ˆz ˆz ˆx ˆx ˆy
Taking the curl again, we get:
A
ˆ 2 f2 ˆ 2 f1 ˆ 2 f1 ˆ 2 f3 ˆ 2 f3 ˆ 2 f2 ˆ 2 f2 ˆ 2 f1
Ò ◊ (Ò ◊ F) = ≠ ≠ + , ≠ ≠ + ,
ˆyˆx ˆy 2 ˆz 2 ˆzˆx ˆzˆy ˆz 2 ˆx2 ˆxˆy
B
ˆ 2 f1 ˆ 2 f3 ˆ 2 f3 ˆ 2 f2
≠ 2
≠ 2
+
ˆxˆz ˆx ˆy ˆyˆz
Putting these two calculations together, and using the fact that partial derivatives
commute by Clairaut’s theorem (since the vector fields are assumed to be smooth), we
get:
Ò(Ò · F) ≠ Ò ◊ (Ò ◊ F)
CHAPTER 4. k-FORMS 143
A B
ˆ 2 f1 ˆ 2 f1 ˆ 2 f1 ˆ 2 f2 ˆ 2 f2 ˆ 2 f2 ˆ 2 f3 ˆ 2 f3 ˆ 2 f3
= + + , + + , + +
ˆx2 ˆy 2 ˆz 2 ˆx2 ˆy 2 ˆz 2 ˆx2 ˆy 2 ˆz 2
ˆ2F ˆ2F ˆ2F
= + + .
ˆx2 ˆy 2 ˆz 2
6. In this problem we prove the statement in Example 4.8.6 about Maxwell’s equations.
(a) Write down the action of the Hodge star operator on basic k-forms in Minkowksi
R4 (see Footnote 4.8.2 ).
(b) Using your result in part (a), show that the two equations
dF =0,
d ı F =J,
We study how two-forms can be integrated along surfaces, which leads us to introduce the
concept of surface integrals (also called “flux integrals”).
Objectives
You should be able to:
144
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 145
In other words, we just evaluate the function at P œ Rn , and multiply by the sign corresponding
to the chosen orientation.
To define the integral over a set of oriented points, we sum up the integrals over each point
separately. ⌃
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 146
For instance, in this language we can define the integral of a function f over two oriented
points {(P0 , ≠), (P1 , +)} as:
⁄
f = f (P1 ) ≠ f (P0 ).
{(P0 ,≠),(P1 ,+)}
And that’s basically it, as far as zero-forms go. There’s no question of parametrization
here, and the integral is clearly oriented by definition. There’s no step 6, as there is no such
thing as an exact zero-form.
Example 5.1.3 Integral of a zero-form at points. Consider the function f (x, y, z) =
xy + z on R3 . Pick the two points P0 = (0, 0, 0) and P1 = (1, 1, 1). Let’s give P0 a negative
orientation, and P1 a positive orientation. Then
⁄
f = f (P1 ) ≠ f (P0 ) = f (1, 1, 1) ≠ f (0, 0, 0) = 2.
{(P0 ,≠),(P1 ,+)}
where on the right-hand-side we use the standard definition of definite integrals from calculus.
⌃
STEP 3. Next step: introduce parametric curves – : [a, b] æ R . This was done in Defini-
n
tion 3.2.1. We do not repeat the definition here, but simply restate that it induces an orientation
on the image curve C = –([a, b]) µ Rn , and on its boundary ˆC = {(–(a), ≠), (–(b), +)} if it
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 147
is not closed. 1 We will write ˆ– to denote the boundary of the parametric curve with its
induced orientation.
STEP 4. We can then define the integral over a parametric curve – using the pullback.
This is Definition 3.3.2.
Definition 5.1.6 (Oriented) line integrals. Let Ê be a one-form on an open subset U of
Rn , and let – : [a, b] æ Rn be a parametric curve whose image C = –([a, b]) µ U . We define
the integral of Ê along – as follows:
⁄ ⁄
Ê= –ú Ê,
– [a,b]
On the right-hand-side, from our definition of integration of zero-forms over oriented points
(Definition 5.1.2), we get: ⁄
f = f (b) ≠ f (a).
ˆ([a,b])
1
This is clear since the domain of a parametric curve is always the interval [a, b] with its canonical orientation.
Since the canonical orientation induces the orientation {(a, ≠), (b, +)} on the boundary of the interval, it
induces the orientation {(–(a), ≠), (–(b), +)} on the boundary of the parametric curve.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 148
which is the statement in Theorem 3.4.1. In particular, if the curve is closed, ˆ– is the empty
set, and the right-hand-side vanishes, as in Corollary 3.4.3.
5.1.4 Exercises
1. Evaluate the integral of the zero-form f (x, y, z) = sin(x) + cos(xyz) + exy at the set of
oriented points S = {(p, +), (q, ≠), (r, +)} with
3 4
fi 2
p = (0, 0, 5), q= , ,fi , r = (fi, 1, 1).
2 fi
since a is a zero of f . Since the result is the same regardless of what orientation the
point a has, we conclude that the integral does not depend on the orientation of the
point. (That’s of course only because the point a is a zero of f , that wouldn’t be true for
an arbitrary point).
3. Suppose that f is a smooth function on Rn , and let p, q œ Rn be two distinct points such
that f (p) = f (q). Show that the line integral of df along any curve starting at p and
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 149
ending at q is zero.
Solution. Let – be any parametric curve whose image starts at p and ends at q. By
the Fundamental Theorem of line integrals, we know that
⁄ ⁄ ⁄
df = f= f = f (q) ≠ f (p) = 0,
– ˆ– {(p,≠),(q,+)}
where the last equality follows since we assume that f (p) = f (q). Therefore, the line
integral of df along any such parametric curve – is zero.
4. Let f be a zero-form on Rn , and S = {(a, +), (≠a, ≠)} for some point a œ Rn (≠a denotes
the point in Rn whose coordinates are minus those of a). Show that
Y
⁄ ]0 if f is even,
f=
S [2f (a) if f is odd.
for some parametric curve – : [a, b] æ R2 . Show that the image curve C = –([a, b]) is
not a closed curve, i.e. it must have boundary points.
Solution. First, we notice that Ê is exact. Indeed, let f (x, y) = x2 y. Then
ˆf ˆf
df = dx + dy = 2xy dx + x2 dy = Ê.
ˆx ˆy
But then, by the Fundamental Theorem of line integrals, we know that
⁄ ⁄ ⁄
Ê= df = f.
– – ˆ–
In particular, if the image curve is closed, then the boundary set is empty, i.e. ˆ– = ÿ,
and the right-hand-side is zero. But the question states that it is non-zero; it is equal to
5. Therefore, the image curve cannot be closed.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 150
6. You want to impress your calculus teacher, and you tell her that “integration by parts”
can be rewritten as the “simple” statement that
⁄ ⁄
d(f g) = f g,
[a,b] ˆ([a,b])
i.e. it is just the Fundamental Theorem of Calculus (part II) for a product of functions
(f, g are differentiable functions on R).
Explain why this is equivalent to integration by parts for definite integrals.
Solution. First, using the graded product rule for the exterior derivative, we know
that d(f g) = gdf + f dg. So we can write the left-hand-side as
⁄ ⁄ ⁄ b ⁄ b
d(f g) = (gdf + f dg) = gdf + f dg.
[a,b] [a,b] a a
As for the right-hand-side, the boundary of the interval is ˆ([a, b]) = {(a, ≠), (b, +)}. So
it can be rewritten as ⁄
f g = f (b)g(b) ≠ f (a)g(a).
ˆ([a,b])
Objectives
You should be able to:
• Determine whether two choices of ordered bases on Rn induce the same or opposite
orientations.
5.2.1 Orientation of Rn
Recall that a choice of orientation on R is a choice of direction: either that of increasing real
numbers, or of decreasing real numbers. We defined the direction of increasing real numbers
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 151
as being the positive orientation, and called it the canonical orientation. We now generalize
this to Rn .
Definition 5.2.1 Orientation of Rn . An orientation of the vector space Rn is determined
by a choice of ordered basis on Rn . We think of the orientation as a “twirl”, which starts at
the first basis vector, then rotates to the second basis vector, to the third, and so on and so
forth. Ordered bases that generate the same twirl correspond to the same orientation of the
vector space. ⌃
What this definition is saying is that to give an orientation to R , we pick a choice of
n
ordered basis. But not all ordered bases give rise to different orientations; if they generate the
same twirl, they correspond to the same orientation of the vector space. If you think about
it carefully (think about R2 and R3 first), you will see that for any Rn , there are only two
distinct choices of twirl, and hence only two distinct choices of orientation.
This is due to a basic fact in linear algebra. Any two ordered bases of Rn are related
by a linear transformation with non-vanishing determinant. One can see that two ordered
bases that are related by a linear transformation with positive determinant generate the same
twirl, while they generate opposite twirl if they are related by a linear transformation with
negative determinant. So we can group oriented bases into two equivalence classes, depending
on whether they are related by linear transformations with positive or negative determinants.
These equivalence classes of oriented bases correspond to the two choices of orientation on Rn .
Next we define the notion of canonical orientation.
Definition 5.2.2 Canonical orientation of Rn . Let (x1 , . . . , xn ) be coordinates on Rn ,
and {e1 , . . . , en } be the ordered basis
with ei the unit vector pointing in the positive xi -direction. We call the orientation described
by this ordered basis the canonical orientation of Rn with coordinates (x1 , . . . , xn ), and
denote it by (Rn , +). We denote by (Rn , ≠) the vector space Rn with the opposite choice of
orientation. ⌃
As always, the definition will be clearer with low-dimensional examples. First, let us show
that we recover our previous definition of orientation in R.
Example 5.2.3 Orientation of R and choice of positive or negative direction. If x
is a coordinate on R, the canonical orientation is specified by the basis vector e1 = 1 in the
positive x-direction. Thus the canonical orientation corresponds to the direction of increasing
real numbers, as mentioned before. We denote it by (R, +).
Another choice of basis in R would be f1 = ≠1. It is related to e1 by the linear transfor-
mation f1 = ≠e1 , which has negative determinant. Thus f1 generates the opposite orientation;
indeed, it points in the direction of decreasing real numbers, which correspond to the orientation
(R, ≠). So we recover our previous definition of orientation for R. ⇤
Example 5.2.4 Orientation of R2 and choice of counterclockwise or clockwise
rotation. Let (x, y) be coordinates on R2 . The canonical orientation is specified by the
ordered basis {e1 , e2 }, with e1 = (1, 0) and e2 = (0, 1) being the unit vectors pointing in
the positive x- and y-directions. In two dimensions, the twirl generated by an ordered basis
corresponds to a choice of direction of rotation, from the first basis vector to the second.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 152
• A boundary point is a point p œ R2 such that all disks centered at p contain points in
D and also points not in D. We define the boundary of D, which we denote by ˆD, to
be the set of all boundary points of the region D.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 153
• We say that a region D is closed if it contains all its boundary points, that is, ˆD ™ D.
We say that it is open if it contains none of its boundary points, that is, D fl ˆD = ÿ.
• We say that a region D is path connected (or connected) if any two points in D can
be connected by a path within D. In other words, it is path connected if it has only one
component.
• We say that a region D is simply connected if it is path connected and all simple
closed curves (loops) in D can be continuously contracted to a point within D. In other
words, it is simply connected if it has only one component and no holes.
⌃
Remark 5.2.7 We will often consider closed, simply connected, bounded regions D µ R2 .
Such a region can be constructed by considering a simple closed curve C µ R2 , and letting the
region D be the closed curve and its interior. The boundary of the region is then ˆD = C, i.e.
the closed curve that we started with. With this construction, it is clear that D is bounded,
and it is simply connected since it has one component and no holes.
We note however that the closed curve C does not have to be a parametric curve; it may
have kinks and corners, that is, it could be piecewise parametric. For instance, C could be a
triangle, or a rectangle, etc. It needs to be simple however, i.e. not have self-intersection.
Next we define the orientation of a closed bounded region D µ R2 . Recall that we defined
the orientation of a closed interval in R as being a choice of direction, just as for R; so we
can think of the orientation of an interval as being induced by a choice of orientation on the
surrounding vector space R. We do the same for regions in R2 .
Definition 5.2.8 Orientation of a closed bounded region in R2 . Let D µ R2 be a
closed bounded region in R2 , and choose an orientation on R2 . We define the orientation of
the region D as being the orientation induced by the surrounding vector space R2 . We write
D+ for the region D µ R2 with the canonical (counterclockwise) orientation on R2 , and D≠
for the region with the opposite (clockwise) orientation. When we write D without specifying
the orientation, we always mean the region D with its canonical orientation. ⌃
2
As a last step, given a closed bounded region D µ R with a choice of orientation, we need
to define the induced orientation on its boundary ˆD. In the one-dimensional case, given a
closed interval [a, b] œ R with canonical orientation, we defined the induced orientation on its
boundary to be {(a, ≠), (b, +)}. We do something similar in two dimensions. In this case, the
boundary ˆD is a curve (which may have more than one components), and hence the induced
orientation should be a choice of direction on each component.
Definition 5.2.9 Induced orientation on the boundary of a region in R2 . Let D µ R2
be a closed bounded region with canonical orientation, and ˆD its boundary. Imagine that
the region D is on the floor and that you are walking on its boundary. We define the induced
orientation on each boundary component as being the direction of travel along the boundary
keeping the region on your left. If D has the opposite orientation, the induced orientation on
the boundary is the opposite direction of travel. ⌃
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 154
D = {(x, y) œ R2 | x2 + y 2 Æ 1},
which is a disk of radius one centered at the origin. The boundary points in D are the points
on the circle x2 + y 2 = 1. Thus ˆD is the circle of radius one centered at the origin. Since it
is included in D, this means that D is closed. It is clearly bounded, as it can be contained
within a disk. It is also simply connected.
We can give D the canonical orientation induced by the standard choice of ordered basis
on R2 , which is the choice of counterclockwise direction of rotation. The induced orientation
on the boundary is the counterclockwise direction of motion along the circle, as this is the
direction that one needs to move along the circle to keep its interior on the left. ⇤
Example 5.2.11 Closed square in R2 . Consider the region
D = {(x, y) œ R2 | ≠ 1 Æ x Æ 1, ≠1 Æ y Æ 1}.
This corresponds to a square and its interior centered at the origin and with side length 2. It
is bounded as it can be contained within a finite disk. Its boundary ˆD is the square itself.
As it is included in D, D is closed. It is also simply connected. If we choose the canonical
orientation on D, the induced orientation on ˆD corresponds to moving counterclockwise
along the square. ⇤
Example 5.2.12 Annulus in R2 . Consider the region
D = {(x, y) œ R2 | 1 Æ x2 + y 2 Æ 2}.
This corresponds to an annulus with inner radius 1 and outer radius 2. D is certainly bounded
as it is contained within a finite disk. Its boundary has two components, ˆD = ˆD1 fi ˆD2 ,
where ˆD1 is the inner circle of radius 1 and ˆD2 is the outer circle of radius 2. While the
region is connected, it is not simply connected, as there is a hole in the middle.
Suppose that we choose the canonical orientation on D. What is the induced orientation
on the boundary? We have to look at the two components separately. First, along the outer
radius ˆD2 , we need to move in the counterclockwise direction to keep the interior of the
annulus on the left. Thus the induced orientation is counterclockwise. However, for the inner
circle ˆD1 , we need to move in the clockwise direction to keep the interior of the annulus on
the left. So the induced orientation on ˆD1 is clockwise. ⇤
Remark 5.2.13 There is another way that one can think of the induced orientation on the
boundary of a region in R2 . It is a little bit more subtle, and while it is not needed at this
stage, it will be useful to generalize to regions in R3 and Rn , so let us mention it here.
Consider first the case where D µ R2 is a closed, simply connected, bounded region.
Then ˆD is a simple closed curve. Since a curve is a one-dimensional subspace of R2 , at a
point p œ ˆD there are two choices of (length one) normal vectors: one that points “inwards”
(towards D), and one that points “outwards” (away from D). For all points p œ ˆD, we pick
the normal vector that points outwards. This defines an orientation on ˆD as follows; start
with the normal vector, and rotate to a tangent vector in a way that reproduces the orientation
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 155
(the twirl) on the ambient space R2 . This defines a choice of tangent vector at all points on
ˆD, which defines the induced direction of motion or orientation on the curve. It is easy to
see in a figure that if the ambient space R2 is equipped with canonical orientation, choosing
the normal vectors pointing outwards corresponds to choosing the direction of motion keeping
the region on the left.
If the region is not simply connected, its boundary may have many components. Do the
same construction for each component, always choosing the normal vectors pointing outwards
(away from D). This will induce the orientation on each component corresponding to the
direction of motion keeping the region on the left.
This construction in terms of normal vectors is more subtle, but it directly generalizes to
closed bounded regions in Rn , which is nice.
5.2.3 Exercises
1. Determine whether the following regions are bounded, closed, connected, and/or simply-
connected.
(a) D = {(x, y) œ R2 | x Ø 2, y Ø 3}.
(a) D extends forever in the positive x and y directions, so it is not bounded (it is not
of finite extent). The boundary points are the points in D with x = 2 or y = 3,
since they are at the edge of the region. All those points are included in D, so D is
closed. It is connected, as it has only one component, and it is simply-connected,
as there is no hole.
(b) D is the unit disk of radius one, without its boundary the circle of radius one. It is
bounded, since it can be contained within a disk (it is of finite extent). It is not
closed, since the boundary of D is the circle of radius one, which is not included in
D. It is connected (one component) and simply-connected (no hole).
(c) D consists in two separate square components. It is bounded, since the two square
components can be contained within a disk (finite extent). It is closed, since the
boundary points are the edges of the squares, which are all included in D. It is
however not connected (two components), and hence also not simply-connected.
(d) D is the region bounded by ellipse centered at the origin. It is bounded (finite
extent), closed (the ellipse itself, which is the boundary, is included in D), connected
(one component), and simply-connected (no hole).
2. Let {e1 , e2 , e3 } be the canonical basis on R3 . Show that the ordered basis {e2 , e3 , e1 }
induces the same orientation on R3 as the canonical basis.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 156
Solution. To show that two ordered bases induce the same orientation, we need to show
that they are related by a linear transformation with positive determinant. If we write
{f1 , f2 , f3 } = {e2 , e3 , e1 } for the ordered basis, we see that it is related to the canonical
basis by the linear transformation (thinking of the basis vectors as column vectors):
Q R
0 0 1
c d
fi = a1 0 0b ei .
0 1 0
Indeed,
Q RQ R Q R
0 0 1 1 0
c dc d c d
a1 0 0b a0b = a1b ,
0 1 0 0 0
Q RQ R Q R
0 0 1 0 0
c dc d c d
a1 0 0b a1b = a0b ,
0 1 0 0 1
Q RQ R Q R
0 0 1 0 1
c dc d c d
a1 0 0b a0b = a0b .
0 1 0 1 0
But Q R
0 0 1
c d
det a1 0 0b = 1,
0 1 0
and hence the two ordered bases induce the same orientation on R3 .
3. Let D be the upper half of a disk of radius one, including its boundary. Suppose that D
is given the canonical orientation. Write its boundary, with the induced orientation, as
an oriented parametric curve.
Solution. The upper half of a disk of radius one is the region of R2 defined by:
D = {(x, y) œ R2 | x2 + y 2 Æ 1, y Ø 0}.
Its boundary ˆD is the upper half of the circle, and the x-axis between x = ≠1 and x = 1.
Since D has canonical orientation, the induced orientation corresponds to walking along
the boundary curve keeping the region to the left, which means going counterclockwise
along the boundary.
To realize ˆD as a parametric curve, we need to split it in two, since there are corners
where the upper half disk meets the x-axis. Let C be the upper half disk, and L be the
part of the x-axis in the boundary. We can parametrize the two curves separately. For C
we take –1 : [0, fi] æ R2 with
This has the correct orientation, as it goes counterclockwise around the circle. As for the
L, we need to parametrize the line y = 0 from x = ≠1 to x = 1. We take –2 : [≠1, 1] æ R2
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 157
with
–2 (t) = (t, 0).
4. Let A be a closed disk of radius four centered at the origin, and B be the region
Let D be the region that consists of all points in A that are not in B. Is D bounded?
Closed? Connected? Simply-connected? What is the boundary of D? And if D is given
the canonical orientation, what is the induced orientation on the boundary ˆD?
Solution. We note that B is the interior of a square of side length two centered at the
origin. It is enclosed within A, which is a disk of radius four. Therefore, D is certainly
bounded, since it is of finite extent.
The boundary points of D consists of all points on its outer boundary, which is the
circle of radius four, and on its inner boundary, which consists on the points on the square
centered at the origin. Thus, it has two separate components. But all the boundary
points are included in D, and hence D is closed.
It is connected, as it has only one component. However, it is not simply connected,
as any closed loop going around the inner missing square cannot be contracted to a point
within D.
What is the induced orientation on the boundary? Let us denote by C1 the outer
boundary consists of the circle of radius four, and C2 the inner boundary consisting of
the square of side length two. Since D has canonical orientation, the induced orientation
on the boundary will be obtained by walking along the boundary components keeping
the region to the left. Along the outer boundary C1 , we will keep the region to the left if
we walk counterclockwise. However, along the inner boundary C2 , we will keep the region
to the left if we walk clockwise. Therefore, the induced orientation is councterclockwise
on C1 and clockwise on C2 .
5. The parametric curve – : [0, fi] æ R2 with
Figure 5.2.14 The parametric curve – : [0, fi] æ R2 with –(t) = (sin(t), sin(2t)).
Suppose that D is the region consisting of the curve and its interior. What should
the orientation of the region D be so that the induced orientation on its boundary is the
same as the orientation of the parametric curve?
Solution. Let us first find the orientation of the parametric curve. The tangent vector
is
T(t) = (cos(t), 2 cos(2t)).
In particular, at the origin, the tangent vector is T(0) = (1, 2). It points upwards and
in the positive x-direction. Therefore, we see that the parametric curve has clockwise
orientation.
The region D is the region consisting of the curve and its interior. If we walk clockwise
along the curve, the region is on our right. This means that we must give D a clockwise
(or negative) orientation if we want the induced orientation on the boundary to be the
same as the orientation of the parametric curve.
Objectives
You should be able to:
where on the right-hand-side we mean the standard double integral from calculus of the
function f over the region D. If D is given the opposite orientation, we define the integral of
Ê over D≠ as: ⁄ ⁄⁄
Ê=≠ f dA.
D≠ D
⌃
Remark 5.3.2 A bit of notation: we will always use double or triple integral signs to denote
the standard double and triple integrals from calculus, while we will use only one integral sign
when we are integrating a differential form.
It is probably worth recalling here how double integrals are defined, from your previous
calculus course. If D is a rectangular domain, that is,
The notation here means that the inner integral is an integral with respect to x (while keeping
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 160
y fixed), while the outer integral is with respect to y.1 We recall Fubini’s theorem, which
states that the order of integration does not matter:
⁄ d⁄ b ⁄ b⁄ d
f dxdy = f dydx,
c a a c
that is, it does not matter whether you integrate in x first and then in y or the other way
around.
To integrate over a more general closed bounded region D, we proceed as follows. Since D
is bounded, we can take it to be inside a rectangular region. We can then extend the function
f to the rectangular region by setting it to zero everywhere outside D. The double integral
over D is then defined to be the integral of the extended function over the rectangular region,
which can be written as an iterated itegral as above.
There are two types of regions that give rise to nice iterated integrals. If D can be written
as follows:
D = {(x, y) œ R2 | x œ [a, b], u(x) Æ y Æ v(x)},
with u, v : R æ R continuous functions, we say that the region is x-supported (or of type I
). In this case, one can show that the double integral can be written as the following iterated
integral:
⁄⁄ ⁄ b ⁄ v(x)
f dA = f dydx,
D a u(x)
where the inner integral is with respect to x (keeping y fixed), while the outer integral is with
respect to y.
If instead D can be written as:
we say that D is y-supported (or of type II). The double integral is then the iterated
integral:
⁄⁄ ⁄ d ⁄ v(y)
f dA = f dxdy,
D c u(y)
with the inner integral being with respect to y (keeping x fixed), and the outer integral with
respect to y.
Note that rectangular regions are particular cases of both x-supported and y-supported
regions. If a region D is either x-supported or y-supported, we say that it is recursively
supported. Most of the regions that we will deal with will be either recursively supported
regions, or regions that can be expressed as unions of recursively supported regions.
Example 5.3.3 Integral of a two-form over a rectangular region with canonical
orientation. Consider the two-form Ê = xy dx · dy on R2 , and the closed bounded region
D = {(x, y) œ R2 | 0 Æ x Æ 2, ≠1 Æ y Æ 3},
1
Note that on the right-hand-side here there is no wedge product: this is not the integral of a two-form, it
is an interated integral in x and y as you have seen in calculus. The inner integral is with respect to x (keeping
y fixed), while the outer integral is with respect to y.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 161
To evaluate the double integral, we used the standard procedure for evaluating double integrals
over rectangular regions as iterated integrals, with the inner integral being with respect to x
(keeping y constant), and the outer integral being with respect to y. ⇤
Example 5.3.4 Integral of a two-form over an x-supported (or type I) region with
canonical orientation. Consider the two-form Ê = xey dx · dy, and the region
where the right-hand-side is the standard double integral of a function in calculus. The subtelty
is that the right-hand-side is not an oriented integral, while the left-hand-side is. Indeed,
suppose for simplicity that D = [a, b] ◊ [c, d] is a rectangular region, as in Example 5.3.3.
Then, we can interpret the right-hand-side as an iterated integral:
⁄⁄ ⁄ d⁄ b
f dA = f dxdy.
D c a
But then, by Fubini’s theorem, we know that we can exchange the order of integration without
issue. That is,
⁄⁄ ⁄ d⁄ b ⁄ b⁄ d
f dA = f dxdy = f dydx.
D c a a c
At first sight, this may appear problematic, as one could be tempted to reinterpret the last
integral as ⁄
f dy · dx,
D+
but since dx · dy = ≠dy · dx, this would be minus the integral we started with! But this is
incorrect. The subtlety is in the choice of orientation.
The key is that in Definition 5.3.1, we started by choosing coordinates (x, y) on R2 , and
then we wrote the one-form Ê = f dx · dy using the basic two-form dx · dy in which the
differentials dx and dy appear in the same order as the coordinates (x, y). This is important,
as dy · dx = ≠dx · dy. So while we can exchange the order of the iterated integrals once we
have written everything in terms of double integrals, when we write the integral in terms of
differential forms, we must use the correct choice of basic two-form dx·dy with the differentials
appearing in the same order as the coordinates of R2 . That’s because integrals of two-forms
are oriented, while double integrals of functions are not.
Note that there is nothing special about the variables x and y. We could name the
coordinates of R2 anything. For instance, if we choose coordinates (u, v) on R2 , then the
integral of a two-form Ê = f du · dv over a region D is equal to the double integral of f over
D (with a positive sign in front) when D is endowed with the canonical orientation described
by the ordered basis {e1 , e2 }, with the basis vectors e1 = (1, 0) and e2 = (0, 1) pointing in the
positive directions of the coordinates u and v respectively.2
• If „ is orientation-preserving, then
⁄ ⁄
„ú Ê = Ê.
D2 D1
• If „ is orientation-reversing, then
⁄ ⁄
„ Ê=≠
ú
Ê.
D2 D1
where on the right-hand-side of each equation we mean the double integral for the recursively
supported regions D1 in the xy-plane and D2 in the uv-plane.
But recall from your previous calculus course that double integrals satisfy a “transforma-
tion formula”, or “change of variables formula”, which is the natural generalization of the
substitution formula for definite integrals. The transformation formula states that
⁄⁄ ⁄⁄
f (x, y) dxdy = f („(u, v))| det J„ |dudv.
D1 D2
Note that there is now an absolute value around the determinant of the Jacobian. Thus, what
this means is that if our transformation is such that det J„ > 0, then | det J„ | = det J„ , and
⁄ ⁄⁄ ⁄ ⁄
Ê= f (x, y) dxdy = f („(u, v))(det J„ ) du · dv = „ú Ê,
D1 D1 D2 D2
This is the statement of the lemma: integrals of two-forms are oriented and reparametrization-
invariant! ⌅
What is particularly nice with the proof of the lemma is that the transformation (or
change of variables) formula for double integrals is simply the statement that
integrals of two-forms over regions in R2 are invariant under orientation-preserving
reparametrizations! Isn’t that cool? It explains why the determinant of the Jacobian
appears; it comes from pulling back the two-form under the change of variables.
The fact that double integrals involve the absolute value of the determinant of the Jacobian,
while our integrals do not (and change signs under orientation-reversing reparametrizations),
is also interesting. As alluded to above, the reason is that our integrals are oriented, while
standard double integrals in calculus are not.
Example 5.3.8 Area of a disk. Consider the basic two-form Ê = dx · dy on R2 . Let us
define the following x-supported domain:
D = {(x, y) œ R2 | x œ [≠1, 1], ≠ 1 ≠ x2 Æ y Æ 1 ≠ x2 }.
It is easy to see that D is a closed disk of radius one centered at the origin. The integral of
the basic two-form Ê over D should give us the area of the disk, namely fi. We calculate:
⁄ ⁄⁄
Ê= dA
D D
⁄ Ô
1 ⁄ 1≠x2
= Ô dydx
≠1 ≠ 1≠x2
⁄ 1 1 2
= 1 ≠ x2 + 1 ≠ x2 dx
≠1
⁄ 1
=2 1 ≠ x2 dx.
≠1
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 165
We could have instead use a change of variables to evaluate this integral: polar coordinates.
Define the map „ : R2 æ R2 with
the map „ : D2 æ D is bijective and invertible in the interior of D2 . The determinant of the
Jacobian is A B
cos(◊) ≠r sin(◊)
det J„ = det = r cos2 (◊) + r sin2 (◊) = r,
sin(◊) r cos(◊)
which is positive for all points in the interior of D2 . Thus „ is an orientation-preserving
reparametrization, so the integral of the pullback „ú Ê = r dr · d◊ over D2 should give us fi
again. Indeed, we get:
⁄ ⁄ 2fi ⁄ 1
„ Ê=
ú
(det J„ ) drd◊
D2 0 0
⁄ 2fi ⁄ 1
= r drd◊
0 0
⁄ 2fi
1
= d◊
2 0
=fi.
Notice how easier the integral was! That’s of course because polar coordinates are well suited
for evaluating integrals over regions that have circular symmetry. ⇤
5.3.3 Exercises
1
1. Evaluate the integral of the two-form Ê = (1+x+y)2
dx · dy over the rectangular region
Figure 5.3.9 The region D is the region bounded by the two curves shown above, the
y-axis, and the x-axis. The green curve is the upper half of the curve x = y 2 , while the
orange line is the line y = 1.
We can describe this region as an x-supported or y-supported region. If we write it
as a y-supported region, we would write 0 Æ y Æ 1, and 0 Æ x Æ y 2 . Thus
1
= sin(1).
3
2 2 2
4. Consider the two-form Ê = e(x +y ) dx · dy, and let D be the disk of radius one with
counterclockwise orientation. Show that
⁄ ⁄ 1
4
Ê = 2fi rer dr.
D 0
and
D2 = {(r, ◊) œ R2 | r œ [0, 1], ◊ œ [0, 2fi]}.
It maps the rectangular region D2 to the unit disk D. The determinant of the Jacobian
of „ is
A B
ˆx ˆx
det J„ = det ˆr
ˆy
ˆ◊
ˆy
ˆr ˆ◊
A B
cos(◊) ≠r sin(◊)
= det
sin(◊) r cos(◊)
=r.
As r œ [0, 1], this is positive on the interior of D2 , and thus we know that the integral
will be invariant under pullback, that is,
⁄ ⁄
„ Ê=ú
Ê.
D2 D
Therefore,
⁄ ⁄
4
Ê= rer dr · d◊
D D2
⁄ 1 ⁄ 2fi
4
= rer d◊dr
0 0
⁄ 1
4
= rer [◊]◊=2fi
◊=0 dr
0
⁄ 1
4
=2fi rer dr.
0
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 169
5. Consider the two-form Ê = xy dx · dy, and let D be the region bounded by the four
curves y = x1 , y = x4 , y = x and y = 4x, with canonical orientation. Evaluate the integral
of Ê over D using the change of variables
3 4
u
(x, y) = ,v .
v
Figure 5.3.10 The region D is the region bounded by the three curves shown above.
The blue curve is the curve y = 1/x, the orange curve is y = 4/x, the green line is y = x,
and the red line is y = 4x.
We could evaluate the integral of Ê over D by splitting the region D into two sub-
regions, and then realizing these sub-regions as x-supported or y-supported. Or, we can
do a change of variables, as specified in the question.
We consider the change of variables „ : D2 æ D with
3 4
u
„(u, v) = ,v .
v
What is the domain D2 in the (u, v)-plane such that „(D2 ) = D? In the variables u and
v, the four curves y = 1/x, y = 4/x, y = x and y = 4x become
v 4v u u
v= , v= , ,v = , v=4 .
u u v v
Those four bounding equations can be rewritten as
Ô Ô
u = 1, u = 4, v = u, v = 2 u.
Here we used the fact that v is positive, since y = v is positive. So we can describe the
region D2 in the (u, v)-plane as the v-supported region
Ô Ô
D2 = {(u, v) œ R2 | u œ [1, 4], u Æ v Æ 2 u}.
Figure 5.3.11 The region D2 in the (u, v)-plane is the region bounded by the four curves
Ô Ô
shown above. The blue curve is the curve v = u, the orange curve is v = 2 u, and the
two grey vertical lines are u = 1 and u = 4.
Next, we calculate the determinant of the Jacobian of „. We get:
A B
ˆx ˆx
det J„ = det ˆu
ˆy
ˆv
ˆy
ˆu ˆv
A B
1
≠ vu2
= det v
0 1
1
= .
v
As v > 0 on D2 , the determinant of the Jacobian is positive on D2 . So we know that the
integral is invariant under pullback:
⁄ ⁄
Ê= „ú Ê.
D D2
„ú Ê =u(det J„ ) du · dv
u
= du · dv.
v
We then integrate:
⁄ ⁄
u
„ú Ê = du · dv
D2 D2 v
⁄ 4 ⁄ 2Ô u
u
= Ô dvdu
1 u v
⁄ 4 Ô
= u [ln(v)]v=Ôu du
v=2 u
1
⁄ 4
! Ô Ô "
= u ln(2 u) ≠ ln( u) du
1
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 171
⁄ 4
= ln(2) u du
31 4
16 1
= ln(2) ≠
2 2
15 ln(2)
= .
2
For fun, let us show that we would get the same thing by evaluating the integral of
Ê = xy dx · dy directly by splitting the region into two recursively supported sub-regions.
Looking at the graph in Figure 5.3.10, one choice is to split D into the two x-supported
regions D1 and D2 , with
We can then evaluate both integrals separately and add them up to get the result. For
D1 , we calculate:
⁄ ⁄ 1 ⁄ 4x
Ê= xy dydx
D1 1/2 1/x
⁄ 1 C Dy=4x
y2
= x dx
1/2 2 y=1/x
⁄ 1 3 4
1
= x 8x2 ≠ dx
1/2 2x2
5 6x=1
1
= 2x4 ≠ ln(x)
2 x=1/2
15 1
= ≠ ln(2).
8 2
As for D2 , we get:
⁄ ⁄ 2 ⁄ 4/x
Ê= xy dydx
D2 1 x
⁄ 1 C Dy=4/x
y2
= x dx
1 2 y=x
⁄ 1 A B
8 x2
= x ≠ dx
1 x2 2
C Dx=2
x4
= 8 ln(x) ≠ ln(x)
8 x=1
15
=8 ln(2) ≠ .
8
Adding those two integrals, we get:
⁄ ⁄ ⁄
Ê= Ê+ Ê
D D1 D2
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 172
15 1 15
= ≠ ln(2) + 8 ln(2) ≠
8 2 8
15 ln(2)
= .
2
This is the same answer that we obtained previously via our change of variables. Great!
Objectives
You should be able to:
– : D æRn
(u, v) ‘æ–(u, v) = (x1 (u, v), . . . , xn (u, v))
such that:
1. – can be extended to a C 1 -function on an open subset U ™ R2 that contains D;
3. If –(u1 , v1 ) = –(u2 , v2 ) for any two distinct (u1 , v1 ), (u2 , v2 ) œ D, then (u1 , v1 ), (u2 , v2 ) œ
ˆD. In other words, – is injective everywhere except possibly on the boundary of D.
The image S = –(D) is a two-dimensional subspace of Rn , which is the surface itself. We say
that the parametric surface is smooth if – can be extended to a smooth function on an open
subset U ™ R2 containing D. ⌃
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 173
This definition is obviously very similar to the definition of parametric curves in Defini-
tion 3.2.1. The three properties play similar roles: together, they ensure that the image surface
S really looks like we expect a surface to look like, that is, some sort of sheet of rubber that may
have been bent, deformed, warped, stretched, what not, but without introducing any puncture
or tear. Property 3 ensures that our parametrization covers the image surface S exactly once
(except possibly for boundary points). Property 2 also ensures that a parametrization induces
a well defined choice of orientation on the image curve S, although this is more subtle than
for parametric curves, as we will see in Section 5.5.
As for parametric curves, we can distinguish between two types of parametric surfaces,
depending on whether the image surface is closed or not. It is a bit more subtle than for
parametric curves, so let us look again at what it means for a parametric curve to be closed.
In Definition 3.2.2, we said that a parametric curve – : [a, b] æ Rn is closed if the image curve
C = –([a, b]) has no endpoints (it is a loop). If it is not closed, then we define the boundary
ˆC = {–(a), –(b)} as containing the endpoints of the image curve.
One way to think about a closed curve is that it is the boundary of a surface. In fact, one
could say that a curve in Rn is closed if and only if it forms the boundary of a surface. If it is
not closed, then it must have endpoints, and those form the boundary of the image curve.
This definition generalizes naturally to surfaces.
Definition 5.4.2 Closed parametric surfaces. Let – : D æ Rn be a parametric surface
with image surface S = –(D) µ Rn . We say that the parametric surface is closed if and only
if the image surface S is the boundary of a solid in Rn . If it is not closed, then it must have
edges: we call the set ˆS consisting of all the points on the edges of S the boundary of the
surface. ⌃
Remark 5.4.3 Given a parametric surface – : D æ Rn , one should not confuse ˆD, the
boundary of the closed bounded region (the domain of –), with ˆS, the boundary of the image
surface. On the one hand, ˆD is never empty, as the domain of – is a closed bounded domain
-- in fact, ˆD is a simple closed curve in R2 as we assume that D is simply connected. On the
other hand, ˆS may or may not be empty, depending on whether the image surface is closed
or not. Thus, generally, –(ˆD) ”= ˆS. However, from the definition of parametric curves it
follows that ˆS ™ –(ˆD); the points on the edges of S can only come from images of points
on the boundary of the domain D.
Example 5.4.4 The graph of a function in R3 . A large number of surfaces in R3 can be
obtained as the graph of a function f (x, y) of two variables:
z = f (x, y).
This defines a surface in R3 . If we choose (x, y) œ D for some closed, bounded, simply
connected domain D, then we can realize the surface as a parametric surface – : D æ R3 with
Assuming that f : R2 æ R is a C 1 function, we can check that this satisfies the properties
of a parametric surface. Property one is automatically satisfied, as f is C 1 . The tangent
vectors are 3 4 3 4
ˆf ˆf
Tu = 1, 0, , Tv = 0, 1, .
ˆu ˆv
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 174
Those are clearly linearly independent, regardless of f . Finally, – is certainly injective on the
interior of D; in fact, it is injective everywhere on D.
Since – is injective everywhere on D, including its boundary, this means that the image
surface is not closed. In this case, its boundary ˆS = –(ˆD) is the image of the boundary of
the domain. Basically, the map – takes the domain D and simply deform it continuously in
the z-direction. ⇤
Surfaces realized as the graph of a function, as in the previous example, are very easy to
study parametrically. But many surfaces do not arise in this way, and may instead be given
by an implicit equation in Rn . Finding a parametrization then becomes more difficult.
Example 5.4.5 The sphere. As a second example, we consider the sphere of fixed radius
R centered at the origin in R3 . Its equation is
x2 + y 2 + z 2 = R2 .
We cannot think of this as the graph of a function as in the previous example, because we
cannot solve for z. So we need to think a bit more to realize it as a parametric surface.
As this is a sphere, it is natural that spherical coordinates may be useful. A point on the
sphere radius one can be written as
(x(◊, „), y(◊, „), z(◊, „)) = (R sin(◊) cos(„), R sin(◊) sin(„), R cos(◊)).
and
–(◊, „) = (R sin(◊) cos(„), R sin(◊) sin(„), R cos(◊)).
We can check that it satisfies the properties of parametric surfaces. First, – is smooth on
R2 , so Property 1 is fine. Second, it is easy to see geometrically that – is injective everywhere,
except at these points:
• For any two „1 , „2 œ [0, 2fi], –(0, „1 ) = –(0, „2 ) = (0, 0, R). All those points are mapped
to the north pole of the sphere.
• For any two „1 , „2 œ [0, 2fi], –(fi, „1 ) = –(fi, „2 ) = (0, 0, ≠R). All those points are
mapped to the south pole of the sphere.
The important point is that these points where – is not injective are all on the boundary of
the rectangular region D. Therefore Property 3 is satisfied.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 175
Are those linearly independent? Since the z-coordinate of T„ is zero, the only place where it
could be a multiple of T◊ is when the z-coordinate of T◊ is also zero, that is ≠R sin(◊) = 0.
This will only occur at ◊ = 0, fi, which are on the boundary of D. So we conclude that the
tangent vectors must be linearly independent on the interior of D. Alternatively, we could
have calculated the cross-product T◊ ◊ T„ , and showed that it does not vanish on the interior
of D.
Finally, we note that the image surface, which is the sphere, is closed, since it is the
boundary of a three-dimensional solid (the ball consisting of the interior of the sphere and its
boundary). ⇤
Example 5.4.6 The cylinder. Consider the lateral surface of a cylinder of fixed radius R
extending in the z-direction. Suppose that we look at the part of the cylinder from z = 0 to
z = 2 (we only look at the lateral surface of the cylinder, we do not include the top and the
bottom). How do we realize it as a parametric surface?
The equation of the cylinder is
x2 + y 2 = R2 ,
with z running from 0 to 2. To parametrize it, we introduce polar coordinates. Then a point
on the cylinder can be written as
and
–(◊, w) = (R cos(◊), R sin(◊), w).
Does it satisfy the properties of a parametric surface? First, – is smooth on R2 , so Property
1 is satisfied. As for Property 2, – is injective except at the points –(0, w) = –(2fi, w) =
(R, 0, w), for all w œ [0, 2]. But those are on the boundary of the rectangular region D, so it is
fine. Finally, the tangent vectors are
bounded domain in the uv-plane. The idea is to consider the horizontal and vertical lines
in the uv-plane that lie on D; those are the lines u = u0 or v = v0 for constant u0 , v0 . The
image of these lines under the map – will be curves on the image surface S: those are called
the grid curves. They help visualize how the parametrization – maps the region D onto the
image surface S.
Example 5.4.7 Grid curves on the sphere. Consider the sphere of Example 5.4.5, which
is realized as the parametric surface – : D æ R3 with
and
–(◊, „) = (R sin(◊) cos(„), R sin(◊) sin(„), R cos(◊)).
The domain D is a rectangular region. Horizontal lines on D correspond to lines with „ = C
for constants C. The image of the horizontal lines would be the grid curves
Since ◊ is the inclination angle, those curves correspond to curves of constant longitude,
starting at the north pole and ending at the south pole.
As for the vertical lines on D, they are given by the lines with ◊ = K for constants K.
The image of the vertical lines would be the grid curves
Those correspond to the curves of constant latitude, going all around the sphere. ⇤
n = Tu ◊ Tv .
We note here that this normal vector is not normalized, i.e. it does not have length one. To
get a normalized vector we would divide by its norm. When we talk about the normalized
normal vector later on, we will use the notation
Tu ◊ Tv
n̂ = ,
|Tu ◊ Tv |
to avoid ambiguity. ⌃
5.4.4 Exercises
1. Realize the part of the plane z = x ≠ 2 that lies inside the cylinder x2 + y 2 = 4 as a
parametric surface.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 178
–(u, v) = (u, v, u ≠ 2)
is a parametrization of the plane. What we need to determine now is what is the region
D in the (u, v)-plane such that –(D) = S, where S is the part of the plane contained
within the cylinder x2 + y 2 = 4. What we can do is find the boundary curve of the surface
S, which consists in the intersection of the plane z = x ≠ 2 and the cylinder x2 + y 2 = 4.
Using our parametrization above –(u, v), we see that x2 + y 2 = 4 corresponds to the
equation u2 + v 2 = 4. In other words, if we define D to be the disk of radius 2 centered
at the origin, then – maps its boundary (the circle of radius two) to the boundary of the
surface S, and the interior to the interior. So this gives an appropriate choice of domain
D. We can describe D as a u-supported region:
D = {(u, v) œ R2 | u œ [≠2, 2], ≠ 4 ≠ u2 Æ v Æ 4 ≠ u2 }.
Then the surface S is realized as the parametric surface – : D æ R3 , with D above and
–(u, v) = (u, v, u ≠ 2).
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 179
Note that there are many other parametrizations that we could have used. For
instance, since the region D is a disk of radius 2, we could have used polar coordinates.
In other words, if we do the change of coordinates u = r cos(◊), v = r sin(◊), the region
becomes
D2 = {(r, ◊) œ R2 | r œ [0, 2], ◊ œ [0, 2fi],
and –2 : D2 æ R3 with
Figure 5.4.11 The surface S is the part of the sphere with radius 2 (in orange) that lies
above the cone in blue.
How can we realize S as a parametric surface?
We start by parametrizing the sphere of radius 2 centered at the origin. We use
spherical coordinates. A parametrization for the sphere is – : D æ R3 , with
and
–(◊, „) = (2 sin(◊) cos(„), 2 sin(◊) sin(„), 2 cos(◊)).
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 180
This is a parametrization of the sphere, but this is not the surface we are
interested in.
We only want to keep the part of the sphere that lies above the cone z = x2 + y 2 . In
other words, we want to restrict the range of the inclination angle ◊ so that it only goes
from 0 to the angle
where the cone intersects the sphere. What is this angle? The cone
has equation z = x2 + y 2 . Using our parametrization above for points on the sphere,
we see that the points on the sphere that also lie on the cone (i.e. at the intersection of
both surfaces) must satisfy
Ò
2 cos(◊) = (2 sin(◊) cos(„))2 + (2 sin(◊) sin(„))2
=2 sin(◊),
where we used the fact that sin(◊) Ø 0 since ◊ œ [0, fi]. Therefore, we must have
tan(◊) = 1.
and
–2 (◊, „) = (2 sin(◊) cos(„), 2 sin(◊) sin(„), 2 cos(◊)).
3. Consider the parametric surface – : D æ R3 with D = {(u, v) œ R2 | u œ [0, 2], v œ [0, 3]}
and
–(u, v) = (u, v 3 + 1, u + v).
(a) Find the tangent vectors Tu , Tv , and the normal vector n.
(b) Find an equation for the tangent plane to the image surface –(D) at the point
(1, 2, 2).
n =Tu ◊ Tv
Q R
i j k
c d
= det a1 0 1 b
0 3v 2 1
=(≠3v 2 , ≠1, 3v 2 ).
(b) To find an equation of the tangent plane at the point (1, 2, 2), we use the point-
normal form for the equation of a plane. First, we see that –(1, 1) = (1, 2, 2), so the point
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 181
is on the image surface –(D) for the values of the parameters (u, v) = (1, 1). The normal
vector to the plane at that point is n(1, 1) = (≠3, ≠1, 3). Then, by the point-normal
form, we know that the equation of the tangent plane is
n · (x ≠ 1, y ≠ 2, z ≠ 2) = (≠3, ≠1, 3) · (x ≠ 1, y ≠ 2, z ≠ 2) = 0.
Evaluating the dot product, we get the equation of the tangent plane:
≠3x ≠ y + 3z = 1.
4. Consider the curve y = x2 with x œ [0, 2]. Find a parametrization for the surface obtained
by rotating the curve about the y-axis.
Solution. First, we sketch the curve y = x2 in the (x, y)-plane inside R3 :
Figure 5.4.12 The curve y = x2 in the (x, y)-plane inside R3 is shown in blue; the
orange line is the axis of rotation, which is the y-axis.
We rotate the curve about the y-axis, which is the orange line. After rotation, we get
the following surface:
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 182
Figure 5.4.13 The surface obtained by rotating the curve y = x2 about the y-axis.
How do we parametrize this surface? Let’s think about it. First, we can parametrize
the curve y = x2 in the (x, y)-plane within R3 , with x œ [0, 2], by „(t) = (t, t2 , 0) with
t œ [0, 2]. What happens if we rotate the curve about the y-axis? For a fixed value
of t, the point with (x, z)-coordinates (t, 0) gets rotated about the y-axis on a circle
with radius t. We can thus parametrize this circle by (x, z) = (t cos(◊), t sin(◊)), with
◊ œ [0, 2fi]. We do that for all values of t œ [0, 2], and we end up with the surface of
revolution. The resulting parametrization is – : D æ R3 with
and
–(t, ◊) = (t cos(◊), t2 , t sin(◊)).
Objectives
You should be able to:
n = Tu ◊ Tv .
The ordering is important here, since this is what gives the direction of rotation. Changing
from a direction of rotation to the opposite one amounts to exchanging the order of the two
basis vectors, which, in turn, sends n to ≠n. So what matters here is the direction of the
normal vector: it either points in one direction or the opposite direction, and this defines the
two choices of orientation on the tangent plane at this point.
So we can think of an orientation of a surface as an assignment of a normal vector at
all points on S in a continuous manner. Since what matters is the direction of the normal
vector, choosing an assignment of a normal vector at all points on S is basically the same as
choosing a side for the surface S. Which raises the following question: are all surfaces in
R3 orientable? That is, do all surfaces in R3 have two sides?
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 184
Perhaps surprisingly, the answer is no! Not all surfaces are orientable. There exists
surfaces that only have one side! We will come back to this in a second. Let us now define
more carefully the notion of an “orientable surface”.
Definition 5.5.1 Orientable surfaces and orientation. Let S µ R3 be a surface. We
say that S is orientable if there exists a continuous function n : S æ R3 which assigns to all
points p œ S (not on the boundary) the normal vector to the tangent plane Tp S. We say that
it is non-orientable otherwise.
If S is orientable, then an orientation is a choice of continuous function n : S æ R3 . It
assigns to all points on S a normal vector, which basically specifies one side of the surface (and
also determines a direction of rotation on the tangent planes). Saying that the assignment of
the normal vector is continuous amounts to saying that you pick the same side all around the
surface — you do not suddenly jump from one side to the other. ⌃
Most surfaces that we encounter in real life, such as spheres, cylinders, planes, tori, etc.
are orientable. But the prototypical example of a non-orientable surface is the Möbius strip.
This is a strange surface, shown in Figure 5.5.2. (It can be constructed by taking a long
rectangle of paper, giving it one half-twist, and then taping back the ends together.) If you
start at a point on the surface, you can pick a side, which amounts to defining a normal vector.
Then you can move around the surface, always picking the same side. But at some point you
will be back at the point you started with, but on the other side! (Try it!) So the assignment
of normal vectors is not continuous, since you if you stopped just before the point you started
with, the normal vector would suddenly have to jump from one side to the other.
What is going on? The point is that, on the Möbius strip, by doing this assignment of
normal vectors, you end up labeling both sides of the strip. The reason is: the strip really
only has one side! Crazy.
Figure 5.5.2 The Möbius strip (By David Benbennick - Own work, CC BY-SA 3.0, https://
commons.wikimedia.org/w/index.php?curid=50359).
The Möbius strip is very cool, but from now on we will always assume that our surfaces
are orientable.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 185
– : D æR3
(u, v) ‘æ–(u, v) = (x(u, v), y(u, v), z(u, v))
be a parametric surface, and assume that the image surface S = –(D) µ R3 is orientable. The
normal vector
n = Tu ◊ Tv ,
with Tu and Tv the two tangent vectors from Definition 5.4.8, naturally induces an orientation
on S.
Note that the order is important here: we must take the cross-product Tu ◊ Tv with the
two vectors in the same order as the coordinates (u, v) on D µ R2 .
Proof. There isn’t much to prove here. We showed in Definition 5.4.8 and Definition 5.4.9
that the normal vector was well defined (and non-zero) for all points that are not in the image
of the boundary of D. So this gives a continuous assignment of a normal vector, and since we
assume that S is orientable, it induces an orientation on S. ⌅
Remark 5.5.4 As for parametric curves, there is another way of thinking about this statement.
We can think of the parametrization – : D æ R3 as not only mapping the region D, but as
also mapping its orientation. When we define parametric surfaces, we always think of the
domain D as being given the canonical orientation (counterclockwise), which is induced by
the canonical basis {e1 , e2 } on R2 , with e1 = (1, 0), e2 = (0, 1). The vector e1 points in the
u-direction, and hence is mapped to Tu , while the vector e2 points in the v-direction and is
mapped to Tv . So the counterclockwise orientation on D induces the orientation given by the
ordered basis {Tu , Tv } on the tangent planes.
Finally, in Definition 5.2.9 we discussed the induced orientation on the boundary curve
ˆD of a closed bounded region D. If D has canonical orientation, the induced orientation of
the boundary is the direction of motion if you walk along the boundary curve keeping the
region on your left. Similarly, a parametric surface induces an orientation on the boundary
ˆS of the image surface S. Since for parametric surfaces we always start with the canonical
orientation on D, we expect the induced orientation on the boundary of the image surface to
be defined similarly, by walking along the boundary curve keeping the surface on your left.
However, things are more subtle here, since the surface is in R3 , so it’s not immediately clear
what it means to walk along the boundary keeping the surface your left... In which direction
should your head be pointing?
For a region D µ R2 , we of course assumed that you were standing up. One way to think
about this is that you had your head in the positive z-direction (if you assume that D is in
the xy-plane embedded in R3 ). We now know that we can think of an orientation as a choice
of normal vector, and the canonical orientation on the xy-plane corresponds to the normal
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 186
vector pointing in the positive z-direction. In other words, we implicitly assumed that your
head was pointing in the direction of the normal vector (or that you were walking on the side
chosen by the orientation). This now generalizes to surfaces in R3 .
Definition 5.5.5 Induced orientation on the boundary of a parametric surface. Let
– : D æ R3 be a parametric surface, with image surface S = –(D) orientable and oriented by
the parametrization. Let ˆS be the boundary (the edges) of the image surface S. The induced
orientation on the boundary curve ˆS is the direction of travel if you keep the region on the
left, with your head in the direction of the normal vector n = Tu ◊ Tv . We denote by ˆ– the
boundary of the image surface with its orientation induced by the parametrization –. ⌃
Example 5.5.6 Upper half-sphere. Let us realize the upper half-sphere of radius R as a
parametric surface, and study its orientation and the induced orientation on its boundary.
The upper half-sphere has equation
x2 + y 2 + z 2 = R2 ,
keeping only the points with z Ø 0. There are many ways that we can parametrize it. For
instance, we can use either Cartesian or spherical coordinates. Let us do it in spherical
coordinates, and leave the Cartesian coordinates parametrization for Exercise 5.5.4.1.
We recall from Example 5.4.5 the parametrization of the sphere of radius R. A parametriza-
tion of the upper half-sphere is obtained in the same way, but restricting the inclination angle
◊ from 0 to fi/2. We get the parametric surface – : D æ R3 , with
5 6
fi
D = {(◊, „) œ R2 | ◊ œ 0, , „ œ [0, 2fi]},
2
and
–(◊, „) = (R sin(◊) cos(„), R sin(◊) sin(„), R cos(◊)).
Now suppose that D has canonical orientation, as always for parametric surfaces. What is
the induced orientation on the upper half-sphere? We calculate the tangent vectors:
The main question is the direction of the normal vector: does it point inwards (towards the
center of the sphere), or outwards? In other words, is the orientation selecting the inner
surface of the upper half-sphere or the outer surface? To find out, we only need to look at
what happens at a given point on the upper half-sphere (we need to pick a point that
!fi "
is not in
the image of the boundary of D). Let’s pick the point with parameters (◊, „) = 4 , 0 . This
Ô
2
is the point 2 R (1, 0, 1). At this point, the normal vector is:
3 4
fi R2
n ,0 = (1, 0, 1) .
4 2
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 187
We thus see that the normal vector points outwards in the radial direction, i.e. away from the
origin. Therefore, the induced orientation on the upper half-sphere it outwards, i.e. the side
of the surface selected is the outer surface.
What about the induced orientation on the boundary? The boundary of the upper half-
sphere is the circle in the xy-plane with radius R (that’s the edge of the upper half-sphere).
! "
In our parametrization, it corresponds to the points with parameters (◊, „) = fi2 , „ for
„ œ [0, 2fi]. The induced orientation should correspond to the direction of motion if we walk
along the circle, with our head in the direction of the normal vector, keeping the surface on
your left. As the normal vector point outwards, we are walking on the circle with our head
pointing away from the origin. If we keep the surface on our left, we end up walking along
the circle counterclockwise in the xy-plane. This is the induced orientation on the boundary
circle. ⇤
„ú – : D2 æ R3
(s, t) ‘æ („ú x(s, t), „ú y(s, t), „ú z(s, t)) = (x(„(s, t)), y(„(s, t)), z(„(s, t)))
If we denote „(s, t) = (u(s, t), v(s, t)), using the chain rule, we get:
3 4 3 4
ˆu ˆx(u, v) ˆy(u, v) ˆz(u, v) ˆv ˆx(u, v) ˆy(u, v) ˆz(u, v)
Vs = , , + , ,
ˆs ˆu ˆu ˆu ˆs ˆv ˆv ˆv
ˆu ˆv
= Tu + Tv ,
ˆs ˆs
and, similarly,
3 4 3 4
ˆu ˆx(u, v) ˆy(u, v) ˆz(u, v) ˆv ˆx(u, v) ˆy(u, v) ˆz(u, v)
Vt = , , + , ,
ˆt ˆu ˆu ˆu ˆt ˆv ˆv ˆv
ˆu ˆv
= Tu + Tv .
ˆt ˆt
Calculating the cross-product, we get:
3 4
ˆu ˆv ˆv ˆu
Vs ◊ Vt = ≠ Tu ◊ Tv
ˆs ˆt ˆs ˆt
=(det J„ )Tu ◊ Tv .
m = (det J„ )n.
Thus, if det J„ > 0, then the normal vectors of both parametrizations have the same sign (they
pick the same side for the image surface) and the induced orientations are the same, while if
det J„ < 0, they have opposite signs (they pick opposite sides) and the induced orientations
are opposite. ⌅
5.5.4 Exercises
1. Redo the parametrization of the upper half-sphere of radius R, as in Example 5.5.6, but
in Cartesian coordinates. (The region D in this case should be a disk of radius R.) Find
the induced orientation on the image surface by calculating the normal vector.
Solution. The upper half-sphere of radius R is the surface in R3 defined by the equation
x2 + y 2 + z 2 = R2
However, we need to specify D. The region should be the disk that is at the bottom
of the upper half-sphere, which is the disk of radius R centered at the origin. So the
region D in the (u, v)-plane should be given by u2 + v 2 Æ R2 . We can rewrite this as a
u-supported region:
D = {(u, v) œ R2 | u œ [≠R, R], ≠ R2 ≠ u2 Æ v Æ R 2 ≠ u2 .
To find the induced orientation, we need to calculate the normal vector. We first find
the tangent vectors:
3 4
ˆ– u
Tu = = 1, 0, ≠ Ô ,
ˆu R 2 ≠ u2 ≠ v 2
3 4
ˆ– v
Tv = = 0, 1, ≠ Ô .
ˆv R 2 ≠ u2 ≠ v 2
The normal vector is given by the cross-product:
n =Tu ◊ Tv
3 4
u v
= Ô ,Ô ,1 .
2 2
R ≠u ≠v 2 R ≠ u2 ≠ v 2
2
which is positive for ◊ œ [0, fi/2]. Therefore, the two parametrizations should induce the
same orientation, as we found.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 190
2. Consider the surface S consisting of the part of the plane z + x + y = 2 that lies inside
the cylinder x2 + y 2 = 1. Find two parametrizations of S that have opposite orientation.
Solution. First, we can parametrize the plane by –(u, v) = (u, v, 2 ≠ u ≠ v). Now we
need to find a region D in the (u, v)-plane such that –(D) is the part of the plane that
lies inside the cylinder x2 + y 2 = 1. We see that if we take the region D to be the disk
u2 + v 2 = 1, then the boundary circle of D is mapped to the boundary of the surface S,
that is, the closed curve at the intersection of the plane and the cylinder. So this is the
region D that we are looking for. We can realize it as a u-supported region. The final
parametrization would be – : D æ R3 with
D = {(u, v) œ R2 | u œ [≠1, 1], ≠ 1 ≠ u2 Æ v Æ 1 ≠ u2 }
and
–(u, v) = (u, v, 2 ≠ u ≠ v).
What is the induced orientation? The tangent vectors are:
which points in the opposite direction, and hence this parametrization induces the
opposite orientation.
3. We have seen in Example 5.4.4 how we can realize the graph of a function f (x, y) as a
parametric surface in R3 .
(a) Show that the natural parametrization of Example 5.4.4 always induces an orien-
tation on the graph of the function with normal vector pointing in the positive
z-direction.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 191
(b) Find another parametrization of the graph of the function that induces the opposite
orientation.
Solution. (a) Let S µ R3 be the surface given by the graph z = f (x, y) of a function
f , for some region D in the (x, y)-plane. We parametrize it as – : D æ R3 with
where D is the same region D but in the (u, v)-plane. To calculate the normal vector,
we calculate the tangent vectors:
3 4 3 4
ˆf ˆf
Tu = 1, 0, , Tv = 0, 1, .
ˆu ˆv
The normal vector is given by the cross-product:
n =Tu ◊ Tv
Q R
i j k
c d
= det a1 0 ˆf ˆu b
0 1 ˆfˆv
3 4
ˆf ˆf
= ≠ ,≠ ,1 .
ˆu ˆv
In particular, we see that it always points in the positive z-direction, regardless of the
details of the function f .
(b) We can do just as in the previous question; we can exchange u and v. We can
write a new parametrization as –2 : D2 æ R3 , with D2 being the same region as D but
with u and v exchanged, and
which is the same vector (with u and v exchanged) but with the opposite sign. This
parametrization therefore induces the opposite orientation to the parametrization in (a).
4. Consider the surface S consisting of the part of the cone z = x2 + y 2 between z = 1
and z = 5, with normal vector pointing outside the cone. The boundary ˆS of S consists
in two separate oriented closed curves. Realize these two curves as parametric curves,
with orientations that are consistent with the induced orientation on the boundary ˆS.
Solution. The cone is shown in the figure below:
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 192
Figure 5.5.8 The part of the cone z = x2 + y 2 between z = 1 and z = 5.
The boundary has two components: the top boundary circle and the bottom boundary
circle. If we call C1 the top boundary circle, and C2 the bottom one, then ˆS = C1 fi C2 .
These two curves are given by the curves:
What is the induced orientation on these boundary curves? The surface of the cone is
oriented with an outwards pointing normal vector. To get the induced orientation on
the boundary curves, we should walk along the curves, keeping the region on our left,
with our head pointing in the direction of the normal vector. We see that C1 should be
oriented clockwise, while C2 should be oriented counterclockwise. We now want to find
parametrization of these two curves that are consistent with the induced orientations.
We start with C1 . This is a circle of radius 5 in the plane z = 5, oriented clockwise.
We can parametrize the circle using polar coordinates, but we want to use sine for x and
cosine for y so that the resulting curve is oriented clockwise. We get the parametrization
–1 : [0, 2fi] æ R3 with
–1 (◊) = (5 sin(◊), 5 cos(◊), 5).
As for C2 , it is the circle of radius 1 in the plane z = 1, oriented counterclockwise.
We use polar coordinates again, but cosine for x and sine for y so that it is oriented
counterclockwise. We get –2 : [0, 2fi] æ R3 with:
Objectives
You should be able to:
• Define the surface integral of a two-form along a parametric surface in R3 , and evaluate
it.
• Rewrite the definition of surface integrals as flux integrals for the associated vector field.
• Show that surface integrals change sign under reparametrizations of the surface that
reverse its orientation.
where the integral on the right-hand-side is defined in Definition 5.3.1, with D given the
canonical orientation (which is consistent with the induced orientation on S). ⌃
What is neat is that by Definition 5.3.1, the integral on the right-hand-side can be written
as a standard double integral from calculus. So using the pullback we reduced the evaluation
of integrals of two-forms over surfaces to standard double integrals!
Example 5.6.2 An example of a surface integral. Consider the two-form Ê =
x dy · dz + z dx · dy. Evaluate its surface integral along the oriented parametric surface
– : D æ R3 with D = {(u, v) œ R2 | u œ [0, 1], v œ [0, 2]} and –(u, v) = (u, uv 2 , v).
We calculate the pullback of the two-form:
1 2 1 2
–ú Ê =u v 2 du + 2uv dv · dv + vdu · v 2 du + 2uv dv
=(uv 2 + 2uv 2 )du · dv
=3uv 2 du · dv.
The surface integral then becomes:
⁄ ⁄
Ê= 3uv 2 du · dv
– D
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 194
⁄ 2⁄ 1
=3 uv 2 dudv
0 0
⁄ 2 C D1
2 u2
=3 v dv
0 2 0
⁄
3 2 2
= v dv
2 0
v 3 --2
= -
2 0
=4.
Note that it is important here that we wrote the two-form –ú Ê in terms of the basic two-form
du · dv (not dv · du) before rewriting as a double integral, as the orientation on D is the
canonical orientation with respect to the coordinates (u, v) on D µ R2 . ⇤
Proof. Let us rewrite the two surface integrals using the pullback. First,
⁄ ⁄
Ê= –ú Ê.
– D1
Second, ⁄ ⁄
Ê= (– ¶ „)ú Ê.
„ú – D2
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 195
From Exercise 4.7.5.7, we know that we can pullback through a chain of maps D2 æ D1 æ R3
„ –
in two different ways, but it gives the same thing: (– ¶ „)ú Ê = „ú (–ú Ê). Therefore,
⁄ ⁄
Ê= „ú (–ú Ê).
„ú – D2
In the end, what we need to compare is two integrals over regions in R2 . But this is precisely
the result of Lemma 5.3.7, for the two-form –ú Ê on D1 . This lemma states that if det J„ > 0,
then ⁄ ⁄
–ú Ê = „ú (–ú Ê),
D1 D2
while if it is orientation-reversing,
⁄ ⁄
Ê=≠ Ê.
– „ú –
⌅
Surface integrals are oriented and reparametrization-invariant, as we want. Nice! As a
result, while we use a parametrization to define a surface integral, the integral can really be
thought of as being defined intrinsically in terms of the surface S and its orientation.
⌅
It is interesting to remark here that while pulling back a one-form along a parametric
curve picked the tangential component of the vector field along the curve, pulling back a
two-form along a parametric surface picks the normal component of the vector field. This
has an interpretation in physics, as we will see. Just as line integrals were related to the
calculation of work (and hence the tangential component of the force field to the direction of
motion was the relevant one), surface integrals are related to the calculation of flux, for which
the normal component of the force field is the relevant one.
With this result, we can rewrite surface integrals directly as double integrals:
Corollary 5.6.5 Let Ê be a two-form on an open subset U ™ R3 , with associated vector field
F. Let – : D æ R3 be a parametric surface whose image surface S = –(D) µ U is orientable,
with –(u, v) = (x(u, v), y(u, v), z(u, v)). Then
⁄ ⁄⁄
Ê= (F(–(u, v)) · n) dA,
– D
where on the right-hand-side this is a double integral over the region D in the uv-plane.
Proof. This follows directly from Lemma 5.6.4, the definition of surface integrals Definition 5.6.1,
and the definition of integrals over region in R2 Definition 5.3.1. We have:
⁄ ⁄
Ê= –ú Ê
– ⁄D
= (F(–(u, v)) · n) du · dv
D
⁄⁄
= (F(–(u, v)) · n) dA.
D
⌅
This is how surface integrals are generally defined in standard vector calculus textbooks.
Remark 5.6.6 Sometimes, the following shorthand notation is used:
⁄⁄ ⁄⁄
F · dS := (F(–(u, v)) · n) dA.
S D
Such integrals are also called flux integrals, because of the physics interpretation, which
we will study in Section 5.9. The integral calculates the flux of the vector field F across the
surface S in the direction of the normal vector n specified by the orientation.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 197
The surface integral then becomes (recall that we need to add a minus sign since the surface
integral is with respect to the inwards orientation, while our parametrization induces the
outwards orientation):
⁄⁄ ⁄ 2fi ⁄ fi
2
F · dS = ≠ sin3 (◊)d◊d„
S 0 0
⁄ 2fi 5 6fi
1 2
=≠ ≠ cos(◊) + cos3 (◊) d„
0 3 0
⁄
2 2fi
=≠ d„
3 0
4fi
=≠ .
3
⇤
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 198
5.6.4 Exercises
1. Find the surface integral of the two-form
Ê = y dy · dz ≠ (x + y) dz · dx
Then we integrate, recalling that we have to add a minus sign because of the orientation:
⁄ ⁄
≠ Ê =≠ –ú Ê
– D
⁄ fi ⁄ 2fi
=8 sin3 (◊) sin2 („)d„d◊
fi/2 0
⁄ fi
=8fi sin3 (◊)d◊
fi/2
16fi
= ,
3
where I used the fact that
⁄ 2fi ⁄ fi
2
sin2 („)d„ = fi, sin3 (◊)d◊ = ,
0 fi/2 3
which you can check independently using trigonometric identities to evaluate the definite
integrals.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 199
–(u, v) = (u + v, u ≠ v, 1 + 2u + v).
Solution. Let us solve this problem using the vector field approach. We know the
parametric surface; we need to calculate the normal vector. The tangent vectors are
(We note that those are constant vectors, since the image surface is planar.) The normal
vector is then
n =Tu ◊ Tv
Q R
i j k
c d
= det a1 1 2 b
1 ≠1 1
=(3, 1, ≠2).
Ê = x dy · dz + z dx · dy
over the surface of the cube with vertices (±1, ±1, ±1), with orientation given by a
normal vector pointing outwards.
Solution. The cube is shown in the figure below. It has six sides, which we consider as
separate parametric surfaces, Si with i = 1, . . . , 6.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 200
Figure 5.6.8 The surface S is the cube shown in the figure. We consider the six sides as
separate parametric surfaces.
All six sides can be parametrized easily, with the same domain D. We write the
parametrizations as –i : D æ R3 , for i = 1, . . . , 6, with
D = {(u, v) œ R2 | u œ [≠1, 1], v œ [≠1, 1]},
and
–1 (u, v) =(1, u, v)
–2 (u, v) =(≠1, u, v)
–3 (u, v) =(u, 1, v)
–4 (u, v) =(u, ≠1, v)
–5 (u, v) =(u, v, 1)
–6 (u, v) =(u, v, ≠1)
Calculating the normal vectors, we see that the normal vectors for –1 , –4 , –5 point
outwards, while the normal vectors for –2 , –3 , –6 point inwards. So we can write our
desired surface integral as
⁄ ⁄ ⁄ ⁄ ⁄ ⁄ ⁄
Ê= Ê≠ Ê≠ Ê+ Ê+ Ê≠ Ê.
S S1 S2 S3 S4 S5 S6
Therefore,
⁄ ⁄ ⁄ ⁄ ⁄ ⁄ ⁄
Ê= Ê≠ Ê≠ Ê+ Ê+ Ê≠ Ê
S S S2 S3 S4 S5 S6
⁄1
=4 du · dv
D
⁄ 1 ⁄ 1
=4 dudv
≠1 ≠1
=16.
4. Let S be the boundary of the region in R3 enclosed by the cylinder x2 + y 2 = 4, the
paraboloid z = x2 + y 2 , and the plane z = 9. Find the surface integral of the vector field
F(x, y, z) = (x2 , y 2 , z 2 )
Figure 5.6.9 The surface S has three components: S1 is the part of the paraboloid (in
blue) with z Æ 4; S2 is the lateral surface of the cylinder (in orange) for 4 Æ z Æ 9; S3 is
the disk x2 + y 2 Æ 4 in the plane z = 9 (in green).
To evaluate the surface integral of Ê over S, we evaluate it over the three components
separately and then add up the integrals.
Let us consider S1 first, which is the part of the paraboloid z = x2 + y 2 with 0 Æ z Æ 4.
We realize it as the parametric surface –1 : D1 æ R3 with
and
–1 (u, ◊) = (u cos(◊), u sin(◊), u2 ).
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 203
We need to make sure that the normal vector points inwards. The tangent vectors are
As u Ø 0, we see that the normal vector points in the positive z-direction, which means
that it points inwards. So we’re good!
We then evaluate the surface integral:
⁄⁄ ⁄ 2 ⁄ 2fi 1 2
F · dS1 = (u2 cos2 (◊), u2 sin2 (◊), u4 ) · (≠2u2 cos(◊), ≠2u2 sin(◊), u) d◊du
S1 0 0
⁄ 2 ⁄ 2fi 1 2
= ≠2u4 (cos3 (◊) + sin3 (◊)) + u5 d◊du
0 0
⁄ 2
=2fi u5 du
0
64fi
= ,
3
s s
where we used the fact that 02fi cos3 (◊)d◊ = 02fi sin3 (◊)d◊ = 0.
Let us now consider S2 , which is the cylinder x2 + y 2 = 4, 4 Æ z Æ 9. We parametrize
it as –2 : D2 æ R3 with
and
–2 (u, ◊) = (2 cos(◊), 2 sin(◊), u).
We look at the normal vector. The tangent vectors are
We thus see that at a point with coordinates (2 cos(◊), 2 sin(◊), u) on the cylinder, the
normal vector is (≠2 cos(◊), ≠2 sin(◊), 0), which points inwards. Good!
We evaluate the surface integral:
⁄⁄ ⁄ 9 ⁄ 2fi 1 2
F · dS2 = (4 cos2 (◊), 4 sin2 (◊), u2 ) · (≠2 cos(◊), ≠2 sin(◊), 0) d◊du
S2 4 0
⁄ 9 ⁄ 2fi 1 2
= ≠8 cos3 (◊) ≠ 8 sin3 (◊) d◊du
4 0
=0,
s 2fi s 2fi
since, as above, 0 cos3 (◊)d◊ = 0 sin3 (◊)d◊ = 0.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 204
and
–3 (u, ◊) = (u sin(◊), u cos(◊), 9).
The tangent vectors are
Finally, we compute the surface integral over S by adding up the three surface
integrals aboves. We get:
⁄⁄ ⁄⁄ ⁄⁄ ⁄⁄
F · dS = F · dS1 + F · dS2 + F · dS3
S S1 S2 S3
64fi
= + 0 ≠ 324fi
3
908fi
=≠ .
3
Done! :-)
5. An “inverse square field” is a vector field F that is inversely proportional to the square
of the distance from the origin. It is very important in physics, as it describes many
physical phenomena, such a gravity, electrostatics, etc. It can be written as
1
F(x, y, z) = C (x, y, z),
(x2 + y2+ z 2 )3/2
for some constant C œ R. Show that the flux (the surface integral) of F across a sphere
S centered at the origin (in the outwards direction) is independent of the radius of S.
This is one of the very nice properties of inverse square fields!
Solution. As usual we use spherical coordinates to parametrize the sphere of fixed
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 205
radius R, as – : D æ R3 , with
and
–(◊, „) = (R sin(◊) cos(„), R sin(◊) sin(„), R cos(◊)).
Let use differential forms to calculate the surface integral. To the vector field F we
associated the two-form
1
Ê=C (x dy · dz + y dz · dx + z dx · dy) .
(x2 + y2+ z 2 )3/2
To calculate the surface integral, we calculate the pullback, noting that x(◊, „)2 +y(◊, „)2 +
z(◊, „)2 = R2 :
C
–ú Ê = (R sin(◊) cos(„)(R cos(◊) sin(„)d◊ + R sin(◊) cos(„)d„) · (≠R sin(◊)d◊)
R3
+ R sin(◊) sin(„)(≠R sin(◊)d◊) · (R cos(◊) cos(„)d◊ ≠ R sin(◊) sin(„)d„)
+ R cos(◊)(R cos(◊) cos(„)d◊) ≠ R sin(◊) sin(„)d„) · (R cos(◊) sin(„)d◊ + R sin(◊) cos(„)d„))
CR3
= (sin3 (◊) cos2 („) + sin3 (◊) sin2 („) + cos2 (◊) sin(◊) cos2 („) + cos2 (◊) sin(◊) sin2 („))d◊ · d„
R3
=C sin(◊)d◊ · d„.
In particular, we see that it does not depend on the radius R of the sphere! The surface
integral is then
⁄ ⁄
Ê= –ú Ê
– D
⁄ 2fi ⁄ fi
=C sin(◊)d◊d„
0 0
⁄ 2fi
=2C d„
0
=4fiC,
Objectives
You should be able to:
• State Green’s theorem for integrals of exact two-forms over closed bounded regions in
R2 .
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 206
• Use Green’s theorem to evaluate integrals of exact two-forms over closed bounded regions
in R2 .
• Use Green’s theorem to evaluate line integrals of one-forms along simple closed curves
in R2 .
Proof. We will only prove Green’s theorem for recursively supported regions. (In fact, we
will only write the proof for x-supported regions, but the proof for y-supported regions is
analogous.) For more general closed bounded regions, one can prove Green’s theorem by
rewriting the region as an union of recursively
1 supported
2 regions.
We write ÷ = f dx + g dy and d÷ = ˆx ≠ ˆy dx · dy.
ˆg ˆf
for some continuous functions u, v : R æ R. The boundary ˆD is a closed simple curve, with
counterclockwise orientation. It can be split into four curves:
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 207
In fact, for C2 and C3 , we will pick the opposite orientations (as it simplifies the parametriza-
tions), and add a negative sign in front of the line integrals. Those four curves can be realized
as parametric curves (C2 and C3 have opposite orientations):
Putting those together, remembering that we need to add a negative sign for C2 and C3 , we
get:
⁄ ⁄ ⁄ ⁄ ⁄
÷= ÷≠ ÷≠ ÷+ ÷
ˆD –1 –2 –3 –4
⁄ 1
= (g(b, u(b) + (v(b) ≠ u(b))t)(v(b) ≠ u(b)) ≠ g(a, u(a) + (v(a) ≠ u(a))t)(v(a) ≠ u(a))) dt
0
⁄ b
! "
+ f (t, u(t)) ≠ f (t, v(t)) + g(t, u(t))uÕ (t) ≠ g(t, v(t))v Õ (t) dt.
a
s
Next, we need to evaluate the integral D d÷. Here we will do our favourite trick: the
pullback. Instead of using the x-supported domain D directly, we will pullback to a rectangular
domain. This will allow us to integrate the two-form. Consider the function „ : D2 æ D,
where
D2 = {(s, t) œ R2 | s œ [a, b], t œ [0, 1]},
with
„(s, t) = (s, u(s) + (v(s) ≠ u(s))t).
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 208
We see that „(D2 ) = D, and that the function is bijective. So by Lemma 5.3.7, we know that
⁄ ⁄
„ú (d÷) = d÷,
D2 D1
so we may as well calculate the left-hand-side. In fact, we also know that „ú (d÷) = d(„ú ÷).
So let us calculate this pullback. First,
„ú ÷ = f („(s, t)) ds + g(„(s, t))(uÕ (s) + (v Õ (s) ≠ uÕ (s))t) ds + g(„(s, t))(v(s) ≠ u(s)) dt.
Then
3
ˆf („(s, t)) ˆ ! "
d(„ú ÷) = ≠ ≠ g(„(s, t))(uÕ (s) + (v Õ (s) ≠ uÕ (s))t)
ˆt ˆt
4
ˆ
+ (g(„(s, t))(v(s) ≠ u(s))) ds · dt.
ˆs
Thus
⁄ ⁄ b⁄ 13 4
ˆf („(s, t)) ˆ ! "
„ (d÷) =
ú
≠ ≠ g(„(s, t))(uÕ (s) + (v Õ (s) ≠ uÕ (s))t) dtds
D2 a 0 ˆt ˆt
⁄ 1⁄ b3 ˆ
4
+ (g(„(s, t))(v(s) ≠ u(s))) dsdt,
0 a ˆs
where we used Fubini’s theorem to exchange the order of two integrals in the second line. We
can then evaluate the inner integrals using the Fundamental Theorem of Calculus, since they
are definite integrals of derivatives. We get:
⁄ ⁄ b
! "
„ (d÷) =
ú
f („(s, 0)) ≠ f („(s, 1)) + g(„(s, 0))uÕ (s) ≠ g(„(s, 1))v Õ (s) ds
D2 a
⁄ 1
+ (g(„(b, t))(v(b) ≠ u(b)) ≠ g(„(a, t))(v(a) ≠ u(a))) dt.
0
Magic: this
s
is exactly the same as the expression that we found many lines above for the line
integral ˆD ÷! Woot woot! Therefore
⁄ ⁄
d÷ = ÷.
D ˆD
⌅
Remark 5.7.2 If D is simply connected, then ˆD is a simple s
closed curve, and D is the
closed curve with its interior. In this case the line integral ˆD ÷ is simply the line integral of
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 209
the one-form ÷ along the simple closed curve with counterclockwise orientation.
However, in the statement of Green’s theorem, D can be more general. For instance, it
could be an annulus.
s
Then ˆD would have two components: the inner and outer circle. The
line integral ˆD ÷ would then be the sum of the line integrals of the one-form ÷ along the two
circles; the outer circle with counterclockwise orientation, and the inner circle with clockwise
orientation, as this is the induced orientation on the boundary according to Definition 5.2.9.
Remark 5.7.3 Following up on the previous remark, we note that we can read Green’s
theorem in two different ways, which is generally the case for all integral theorems of vector
calculus. We could say:
1. The integral of the exact two-form d÷ over the region D µ R2 with canonical orientation
is equal to the integral of the one-form ÷ over the boundary ˆD with the induced
orientation.
2. The integral of the one-form ÷ over the simple closed curve ˆD µ R2 with canonical
orientation is equal to the integral of its exterior derivative d÷ over the interior D with
canonical orientation.
This is just two different readings of the same statement, depending on whether you start on
the left-hand-side or the right-hand-side. Consequently, we can use Green’s theorem either to
evaluate integrals of exact two-forms via reading 1, or to evaluate line integrals over simple
closed curves via reading 2.
In practice however, Green’s theorem is generally useful mostly to evaluate line integrals
by transforming them into surface integrals.
Example 5.7.4 Using Green’s theorem to calculate line integrals. Find the
line integral of the one-form Ê = xy dx + (x + y) dy over the rectangle with vertices
(≠2, ≠1), (2, ≠1), (2, 0), (≠2, 0), with a clockwise orientation.
We could evaluate the line integral using previous techniques, by rewriting the curve as
four parametric curves for each line segment and then use the definition of line integrals. But
let us instead use Green’s theorem.
Let us denote the rectangular curve by C. It is the boundary of the region D consisting of
the interior of the rectangle with its boundary:
Thus we can use Green’s theorem to rewrite the line integral along C as a surface integral
over D.
We have to be careful with orientation though. If D has canonical orientation, the induced
orientation on the boundary rectangle C will be counterclockwise. But the question is asking
us to evaluate the line integral along C with clockwise orientation. So in Green’s theorem,
if we relate the line integral to the surface integral over D with canonical orientation, we
will need to introduce a minus sign. More precisely, let us denote C≠ to be the rectangle
with clockwise orientation, C+ the rectangle with counterclockwise orientation, and D the
rectangular region with canonical (counterclockwise) orientation. Green’s theorem states:
⁄ ⁄ ⁄
Ê=≠ Ê=≠ dÊ.
C≠ C+ D
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 210
s
So instead of calculating the line integral in the problem, we can calculate ≠ D dÊ.
The two-form dÊ is:
It is a good exercise to check that this is the correct answer by evaluating the line integral
using the standard approach with parametric curves. ⇤
While Green’s theorem is mostly useful to evaluate line integrals, we can also use Green’s
theorem in the reverse direction, to evaluate integrals of exact two-forms. Here is an example.
Example 5.7.5 Area of an ellipse. Find the area enclosed by the ellipse
x2 y 2
+ 2 = 1.
a2 b
The area is given by integrating the basic two-form Ê = dx · dy over the region D bounded
by the ellipse with canonical orientation. We could calculate this integral directly, as a double
integral. Or we can use Green’s theorem, if Ê is exact. But it certainly is: for instance, we
can write Ê = d÷ with ÷ = 12 (x dy ≠ y dx) (we could choose other one-forms such that d÷ = Ê
as well, but this one gives a nice and easy integral). Then Green’s theorem states:
⁄ ⁄
Ê= ÷,
D C
where on the right-hand-side we are evaluating the line integral of ÷ = 12 (x dy ≠ y dx) along
the ellipse with counterclockwise orientation.
We can parametrize the ellipse as – : [0, 2fi] æ R2 with –(t) = (a cos(t), b sin(t)). The
orientation is counterclockwise as required. The pullback –ú ÷ is:
3 4
ab 2 ab ab
– ÷=
ú
cos (t) ≠ sin(t)(≠ sin(t)) dt = dt.
2 2 2
The line integral is then
⁄ ⁄ 2fi
ab
÷= dt
C 0 2
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 211
=fiab,
5.7.3 Exercises
1. Evaluate the line integral of the one-form
Ê = ey dx + ex dy
along the rectangle with vertices (0, 0), (3, 0), (3, 2), (0, 2), counterclockwise, using two
different methods: (a) directly, (b) using Green’s theorem.
Solution. (a) Let us call ˆD the path along the rectangle counterclockwise. It is shown
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 212
in the figure below. Figure 5.7.6 The path ˆD along the rectangle.
To evaluate the line integral directly, we need to parametrize separately the four line
segments. We write the parametrizations as –i : [0, 1] æ R, i = 1, . . . , 4, with
–1ú Ê =3 dt,
–2ú Ê =2e3 dt,
–3ú Ê = ≠ 3e2 dt,
–4ú Ê = ≠ 2 dt.
(b) We now use Green’s theorem to calculate the line integral. By Green’s theorem,
we know that ⁄ ⁄
Ê= dÊ,
ˆD D
where D is the interior of the rectangle with its boundary, that is,
dÊ =ey dy · dx + ex dx · dy
=(ex ≠ ey ) dx · dy.
=1 + 2e3 ≠ 3e2 .
Therefore, ⁄
Ê = 1 + 2e3 ≠ 3e2 ,
ˆD
which is the same answer as in part (a), as it should.
2. Use Green’s theorem to evaluate the line integral of the vector field F(x, y) = (x1/3 +
y, x4/3 ≠ y 1/3 ) along the triangle with vertices (1, 3), (2, 2), (1, 1), clockwise.
Solution. Let us call ˆD the path around the triangle clockwise. It is shown in the
figure below.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 214
while moving an object first along the parabola y = x2 from the origin to the point
(2, 4), then along a horizontal line back to the y-axis, and then back to the origin along a
vertical line.
Solution. We denote the closed path by ˆD. It is shown in the figure below.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 216
with D being the region enclosed by the path (with its boundary), with canonical
orientation. The region D can be described as an x-supported region:
dÊ =2y dy · dx + y dx · dy
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 217
= ≠ y dx · dy
where in the last line –i is a parametrization of the line segment Li . More explicitly,
since the line segment Li joins the points (xi , yi ) and (xi+1 , yi+1 ), we can parametrize it
as –i : [0, 1] æ R2 with
Then
–iú ÷ = (xi + t(xi+1 ≠ xi ))(yi+1 ≠ yi ) dt.
The line integral along –i is then:
⁄ ⁄ 1
÷= (xi + t(xi+1 ≠ xi ))(yi+1 ≠ yi ) dt
–i 0
3 4
1
=(yi+1 ≠ yi ) xi + (xi+1 ≠ xi )
2
1
= (yi+1 ≠ yi )(xi + xi+1 )
2
1
= (xi yi+1 ≠ xi+1 yi + xi+1 yi+1 ≠ xi yi ).
2
Finally, we add up these line integrals to get the area of the polygon. We get:
n ⁄
ÿ
A= ÷
i=1 –i
1 ÿn
= (xi yi+1 ≠ xi+1 yi + xi+1 yi+1 ≠ xi yi )
2 i=1
1
= ((x1 y2 ≠ x2 y1 + x2 y2 ≠ x1 y1 ) + (x2 y3 ≠ x3 y2 + x3 y3 ≠ x2 y2 ) + . . .
2
+ (xn y1 ≠ x1 yn + x1 y1 ≠ xn yn )),
where in the last line we used the fact that xn+1 = x1 and yn+1 = y1 . We see that the
terms x1 y1 , x2 y2 , . . . all cancel out, and we are left with
1
A = ((x1 y2 ≠ x2 y1 ) + (x2 y3 ≠ x3 y2 ) + . . . + (xn y1 ≠ x1 yn )).
2
Bingo!
5. We already studied the one-form
y x
÷=≠ dx + 2 dy,
x2 +y 2 x + y2
defined on U = R2 \ {(0, 0)}, since it is the typical example of a one-form that is closed
but not exact. For instance, we proved in Exercise 3.4.3.2 that it is not exact by showing
that its line integral along a circle of radius one around the origin is non-vanishing. In
this problem we show that the line integral of ÷ is in fact non-vanishing for every simple
closed curve that encloses the origin, and vanishing for every simple closed curve that
does not pass through or enclose the origin.
(a) Consider an arbitrary simple closed curve C0 with canonical orientation that does
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 219
not pass through or enclose the origin. Use Green’s theorem to show that
⁄
÷ = 0.
C0
(b) Let C1 be an arbitrary simple closed curve with canonical orientation that encloses
the origin. Explain why the argument of (a) does not apply here. So we need to do
something else to study the line integral of Ê along C1 .
(c) As in (b), let C1 be an arbitrary simple closed curve with canonical orientation
that encloses the origin. Suppose that K is a circle centered at the origin, with a
radius small enough that K lies completely inside C1 . Give K a counterclockwise
orientation. Use Green’s theorem to show that
⁄ ⁄
÷= ÷.
C1 K
You have showed that the line integral of ÷ along an arbitrary simple closed curve that
encloses the origin is non-vanishing (and is in fact equal to 2fi), while the line integral
along an arbitrary simple closed curve that does not pass through or enclose the origin is
vanishing. Isn’t it neat?
Solution. The key in this problem is to be very careful with the domain of definition
of the one-form ÷. The one-form is define on the open set U = R2 \ {(0, 0)}.
As a starting point, recall that the one-form ÷ is closed, that is, dÊ = 0. Indeed,
A B A B
x2 + y 2 ≠ 2y 2 x2 + y 2 ≠ 2x2
d÷ = ≠ dy · dx + dx · dy
(x2 + y 2 )2 (x2 + y 2 )2
(x2 ≠ y 2 ) + (y 2 ≠ x2 )
= dx · dy
(x2 + y 2 )2
=0.
since d÷ = 0.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 220
(b) If C1 is a simple closed curve that encloses the origin, we cannot apply the
argument in (a). The reason is that, if we denote by D1 the region consistin of the closed
curve C1 and its interior, then D1 is not a subset of U , since D1 contains the origin. In
other words, the one-form ÷ is not defined on D1 , which is one of the assumptions in the
statement of Green’s theorem. Therefore, Green’s theorem does not apply. (Indeed, as
we will see, if Green’s theorem applied, it would imply that the line integral of ÷ along
C1 vanishes, which is not true, as we will see.)
(c) We suppose that C1 is an arbitrary simple closed curve with canonical orientation
that encloses the origin, and that K is a circle centered at the origin that lies completely
inside C1 , also oriented counterclockwise. For instance, C1 could be the orange curve in
the figure below, and K the blue circle.
where we used the fact that line integrals pick a sign if we change the orientation. But
the left-hand-side of the above equation is necessarily zero, since d÷ = 0. Therefore,
⁄ ⁄
÷= ÷,
(C1 )+ K+
which is the statement that we were trying to prove, with both curves canonically oriented.
(d) The power of what we did in (c) is to replace the evaluation of the line integral of
÷ along an arbitrary simple closed curve C1 that encloses the origin by the evaluation of
the line integral of ÷ along a circle K centered at the origin of a certain non-zero radius.
But we can evaluate the latter explicitly! Indeed, if K has radius R, for some R œ R>0 ,
then we can parametrize K by – : [0, 2fi] æ R2 with
The pullback of ÷ is
1
–ú ÷ = (≠R sin(◊)(≠R sin(◊)) + R cos(◊)(R cos(◊))) d◊
R2
=d◊.
Therefore,
⁄ ⁄
÷= –ú ÷
– [0,2fi]
⁄ 2fi
= d◊
0
=2fi.
So the line integral of ÷ along any circle K centered at the origin is always non-zero and
equal to 2fi. By (c), we thus conclude that
⁄
÷ = 2fi
C1
for arbitrary simple closed curves C1 that enclose the origin. Very powerful result!
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 222
Objectives
You should be able to:
• State Stokes’ theorem for integrals of exact two-forms over parametric surfaces in R3 .
• Show that Stokes’ theorem implies that the integral of an exact two-form over a closed
surface in R3 vanishes.
• Use Stokes’ theorem to evaluate line integrals of one-forms over closed curves in R3 .
• Use Stokes’ theorem and its consequences to show that a given two-form cannot be
exact.
which is the Fundamental Theorem of line integrals rewritten using integrals of zero-forms.
The integral on the left-hand-side is a line integral along the oriented parametric curve –,
while on the right-hand-side we have the integral of the zero-form f over the oriented boundary
ˆ– of the parametric curve –.
Looking back at Theorem 3.4.1, the proof of the Fundamental Theorem of line integrals
basically amounted to pulling back to the interval [a, b] using the definition of line integrals,
and then using the Fundamental Theorem of Calculus. We now do the same thing for surface
integrals: we pull back to the region D and then use Green’s theorem. The result is known as
Stokes’ theorem.
Theorem 5.8.1 Stokes’ theorem. Let ÷ be a one-form on an open subset U ™ R3 , and
– : D æ R3 a parametric surface whose image S = –(D) µ U is orientable and oriented by
the parametrization. Let ˆ– be the boundary of the surface S with its induced orientation, as
in Definition 5.5.5 (if the image surface is closed, ˆ– is the empty set). Then the integral of
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 223
the exact two-form d÷ along the parametric surface – can be evaluated as:
⁄ ⁄
d÷ = ÷.
– ˆ–
In other words, the surface integral of the exact two-form Ê = d÷ along the parametric surface
– is equal to the line integral of the one-form ÷ along its oriented boundary curve.
Proof. Stokes’ theorem pretty much follows from Green’s theorem by pullback. More precisely,
by definition of the surface integral,
⁄ ⁄
d÷ = –ú (d÷).
– D
We know that the exterior derivative commutes with pullback, see Lemma 4.7.4: –ú (d÷) =
d(–ú ÷). So we can rewrite the surface integral as:
⁄ ⁄
d÷ = d(–ú ÷).
– D
This is now the integral of the exact two-form d(–ú ÷) over a closed bounded region D. By
Green’s theorem, Theorem 5.7.1, we know that
⁄ ⁄
d(–ú ÷) = –ú ÷,
D ˆD
where on the right-hand-side we have a line integral of the one-form –ú ÷ along the boundary
of the closed region D µ R2 .
We are almost done, but not quite: the last step is actually quite subtle. We would like to
say that ⁄ ⁄
?
–ú ÷ = ÷.
ˆD ˆ–
If the parametrization – is injective everywhere, then it maps the boundary ˆD one-to-one to
the boundary ˆS, and so this equality follows from the definition of line integrals, thinking
of – restricted to the boundary ˆD as a parametric curve (or a union of parametric curves).
However, in general, it is a little more subtle; as we remarked in Remark 5.4.3, ˆS ™ –(ˆD),
but the two are not necessarily equal.1 The general proof is beyond the scope of these notes.
In any case, the conclusion would be that
⁄ ⁄
d÷ = ÷,
– ˆ–
to path independence for line integrals of exact one-forms). Second, the integral of an exact
two-form along a closed surface is zero (this statement for surface integrals is analogous to
the statement that the line integral of an exact one-form along a closed curve vanishes). We
state those as the following corollaries.
Corollary 5.8.2 The surface integrals of an exact two-form along two surfaces
that share the same oriented boundary are equal. Let ÷ be a one-form on U ™ R3 .
If – : D1 æ R3 and — : D2 æ R3 are two parametric surfaces whose image surfaces
S1 = –(D1 ) µ U and S2 = —(D2 ) µ U share the same boundary ˆS1 = ˆS2 , with the same
induced orientation, then ⁄ ⁄
d÷ = d÷.
– —
Corollary 5.8.3 The surface integral of an exact two-form along a closed surface
vanishes. Let ÷ be a one-form on U ™ R3 . If – : D æ R3 is a parametric surface whose
image surface S = –(D) µ U is closed (that is, it has no boundary, which means that it is
itself the boundary of a volume), then
⁄
d÷ = 0.
–
Remark 5.8.4 We can read Stokes’ theorem in two ways, just like Green’s theorem.
2. The integral of a one-form ÷ over a closed oriented simple curve ˆS in R3 is equal to the
surface integral of its exterior derivative d÷ over any surface S µ R3 whose boundary is
ˆS, with the appropriate orientation.
These two different readings highlight the two different types of applications of Stokes’ theorem:
either to calculate surface integrals via the first reading, or to calculate line integrals via the
second reading. We note that Corollary 5.8.2 can also be useful computationally, to replace a
complicated surface integral by an easier one.
Example 5.8.5 Using Stokes’ theorem to evaluate a surface integral by transforming
it into a line integral. Use Stokes’ theorem to evaluate the surface integral of the two-form
Ê = 2z dy · dz + dx · dy along the part of the sphere x2 + y 2 + z 2 = 25 that is above the
plane z = 3, with orientation given by an outwards pointing normal vector.
The surface is shown in the following figure:
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 225
Figure 5.8.6 The surface is part of the sphere of radius 5 above the plane z = 3.
To start with, for Stokes’ theorem to apply, the two-form Ê = 2z dy · dz + dx · dy must
be exact. But it easy to see that
÷ = (x ≠ z 2 ) dy ∆ d÷ = Ê,
To evaluate the line integral on the right-hand-side we parametrize the oriented boundary
with – : [0, 2fi] æ R3 ,
–(◊) = (4 cos(◊), 4 sin(◊), 3).
We calculate the pullback of the one-form ÷:
⇤
Example 5.8.7 Using Stokes’ theorem to evaluate a surface integral by using a
simpler surface. Let us consider the same integral as in the previous example, Example 5.8.5.
That is, we want to evaluate the surface integral of the two-form Ê = 2z dy · dz + dx · dy
along the part of the sphere x2 + y 2 + z 2 = 25 that is above the plane z = 3, with orientation
given by an outwards pointing normal vector.
We can use Stokes’ theorem in a different way to evaluate this surface integral. Since we
know that the two-form is exact (as Ê = d÷ with ÷ = (x ≠ z 2 ) du), we know that its surface
integral along any two surfaces S and S Õ that share the same oriented boundary are equal. So
instead of integrating over the part of the sphere above the z = 3 plane, we could replace this
surface by a simpler surface that shares the same boundary, which in this case is the circle
x2 + y 2 = 16 in the plane z = 3, as we saw in Example 5.8.5. In particular, we could take the
surface S Õ to be the disk x2 + y 2 Æ 16 in the plane z = 3, with orientation given by a normal
vector pointing upwards. By Stokes’ theorem, we know that
⁄ ⁄
Ê= Ê,
S SÕ
so we can evaluate the surface integral over the oriented disk instead.
To do so, we parametrize the disk by – : D æ R3 , with
D = {(r, ◊) œ R2 | r œ [0, 4], ◊ œ [0, 2fi]}
and
–(r, ◊) = (r cos(◊), r sin(◊), 3).
It is easy to see that the normal vector n = Tr ◊ T◊ points upwards, as required. We calculate
the pullback two-form:
–ú Ê =2(3)(sin(◊) dr + r cos(◊) d◊) · (0) + (cos(◊) dr ≠ r sin(◊) d◊) · (sin(◊) dr + r cos(◊) d◊)
=r dr · d◊.
The surface integral is then:
⁄ ⁄
Ê= –ú Ê
– D
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 227
⁄ 2fi ⁄ 4
= r drd◊
0 0
⁄ 2fi C 2 Dr=4
r
= d◊
0 2 r=0
⁄ 2fi
=8 d◊
0
=16fi.
Fortunately, this is the same answer that we found in Example 5.8.5, as it should! ⇤
Example 5.8.8 Using Stokes’ theorem to evaluate a line integral by transforming
it into a surface integral. Use Stokes’ theorem to evaluate the line integral of the one-form
Ê = xy dx + yz dy + z 4 dz over the curve C that forms the boundary of the part of the
paraboloid z = 1 ≠ x2 ≠ y 2 with x Ø 0, y Ø 0, z Ø 0. Assume that C is given the orientation
as moving from the x-axis to the y-axis to the z-axis.
It may be hard to visualize the curve and the region at first. A picture is worth a thousand
words. Here is a figure representing the part of the paraboloid z = 1 ≠ x2 ≠ y 2 with x Ø 0,
y Ø 0, z Ø 0.
Figure 5.8.9 The part of the paraboloid z = 1 ≠ x2 ≠ y 2 with x, y, z Ø 0. The curve is the
boundary of the region.
We want to evaluate the line integral of Ê along the boundary C of this region. We could
evaluate the line integral directly. Or we can use Stokes’ theorem, which says that
⁄ ⁄
Ê= dÊ,
C S
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 228
where S is the part of the paraboloid described above. We need to be careful with orientation
here. We assume that C is oriented going from x to y to z. Looking at the figure, we see that
if we walk along the boundary curve in this direction, to keep the surface on the left we must
have our heads pointing away from the origin. So the surface S should be oriented with a
normal vector pointing in that direction.
To evaluate the surface integral of dÊ along S, we first calculate the exterior derivative.
We get:
dÊ = xdy · dx + ydz · dy.
Next, we parametrize the surface S as – : D æ R3 , with
and
–(r, ◊) = (r cos(◊), r sin(◊), 1 ≠ r2 ).
The region D was chosen such that x(r, ◊), y(r, ◊), z(r, ◊) Ø 0 on D. Thus it covers the part of
the paraboloid that we are interested in.
The normal vector is
n =Tr ◊ T◊
=(cos(◊), sin(◊), ≠2r) ◊ (≠r sin(◊), r cos(◊), 0)
=(2r2 cos(◊), 2r2 sin(◊), r).
As its z-component is positive, we see that it points away from the origin, as it should. So the
orientation induced by the parametrization is the correct one.
Finally, we calculate the pullback of the two-form dÊ:
• The surface integral of the curl of a vector field is the same for any two surfaces that
share the same oriented boundary.
• The surface integral of the curl of a vector field along a closed surface always vanishes.
As before, we can use Stokes’ theorem in both directions. For instance, if we are interested
in calculating the line integral of a vector field along a closed curve in R3 (say, to calculate
the work done by a force field on an object moving along the closed curve), then we can use
Stokes’ theorem, which tells us that that the line integral is equal to the surface integral of
the curl of the vector field on any surface whose boundary is the closed curve. In particular,
we are free to choose a surface that is simple enough so that the resulting surface integral is
easy to evaluate.
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 230
5.8.3 Exercises
1. Let S be the part of the cone z = 2 x2 + y 2 below the plane x + z = 1, oriented with
an upward
normal vector. Let S Õ be the part of the plane x + z = 1 contained within the
cone z = 2 x2 + y 2 , also oriented with an upward normal vector. Let ÷ be the one-form
2 3
÷ = ex dx + ey dy + 5(x + z ≠ 1)2 dz.
Solution. First,
the oriented boundary of S is the closed curve at the intersection of
the cone z = 2 x2 + y 2 and the plane x + z = 1, oriented counterclockwise (looking from
above). We notice that the boundary of S Õ is exactly the same. Therefore, by Stokes’
theorem, we know that ⁄ ⁄
d÷ = d÷.
S SÕ
So we may as well evaluate the integral of d÷ along the surface S Õ , which lies within the
plane x + z = 1.
To do so, we need to evaluate the exterior derivative d÷. We get:
2 3
d÷ =d(ex ) · dx + d(ey ) · dy + 5d((x + z ≠ 1)2 ) · dz
=10(x + z ≠ 1) dx · dz.
s
To evaluate the surface integral S Õ d÷, we would need to parametrize the surface S Õ ,
pullback d÷, and evaluate the integral as a double integral. However, we can show that
the integral will be zero right away without going through the whole calculation. Indeed,
when we pullback d÷ by the parametrization, we need to evaluate its component function
on the surface. But the surface S Õ lies within the plane x + z = 1, and we see right away
that the component function vanishes when x + z ≠ 1. In other words, the two-form
d÷ = 0 when restricted to the plane x + z = 1. So it will certainly vanish on S Õ . Therefore,
we can conclude right away that ⁄
d÷ = 0,
SÕ
and, by Stokes’ theorem, ⁄
d÷ = 0.
S
2. Consider the force field
Use Stokes’ theorem to find the work done by the force field when moving an object
counterclockwise along the circle x2 + y 2 = 4 in the plane z = 10.
Solution. Let ˆS be the circle x2 + y 2 = 4 in the plane z = 10 with counterclockwise
orientation, and let S be the disk x2 + y 2 Æ 4 within the plane z = 10, with orientation
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 231
So instead of evaluating the line integral of F along ˆS, we can evaluate the surface
integral of the curl of F along S.
First, we calculate the curl of F. We get:
Ò ◊ F = (xexy , 1 ≠ yexy , y 2 + x2 ).
To calculate the surface integral, we need to parametrize the surface S. We use the
following parametrization: – : D æ R3 , with
with
–(r, ◊) = (r cos(◊), r sin(◊), 10).
The tangent vectors are
Therefore, by Stokes’ theorem we conclude that the work done by the force field when
moving an object counterclockwise along the circle ˆS is 8fi.
3. Let S be the surface consisting of the top and the four sides (but not the bottom) of the
cube with vertices (±1, ±1, ±1) in R3 , with orientation given by an outward pointing
normal vector, and F the vector field
Solution. We can use Stokes’ theorem in two different ways here: either to rewrite the
surface integral as a line integral, or to rewrite it as a surface integral over a simpler
surface. For completeness I will do it both ways.
First, we can use Stokes’ theorem to rewrite the surface integral as a line integral.
Let us do this first. We notice that the boundary curve ˆS of the surface S is the square
with vertices (±1, ±1, ≠1), since the bottom side of the cube is not included in S. The
induced orientation on the boundary amounts to moving counterclockwise along the
square, as seen from above. Thus, by Stokes’ theorem, we know that
⁄⁄ ⁄
(Ò ◊ F) · dS = F · dr,
S ˆS
where the right-hand-side is the line integral of the vector field F along the square with
counterclockwise orientation.
To evaluate the line integral, we split the square into four line segments Li , i = 1, . . . , 4.
We parametrize the line segments by –i : [0, 1] æ R3 with
T1 = (2, 0, 0),
T2 = (0, 2, 0),
T3 = (≠2, 0, 0),
T4 = (0, ≠2, 0).
Adding up these four line integrals, we get ≠4. We conclude, by Stokes’ theorem, that
the surface integral ⁄⁄
(Ò ◊ F) · dS = ≠4.
S
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 233
Another approach would have been to use Stokes’ theorem to rewrite the surface
integral as a simpler surface integral. For instance we could have replaced the surface S
by the missing side of the cube (namely the bottom), as it shares the same boundary
with S. Let us call the bottom side of the cube S Õ . If we give it an upward pointing
normal vector, it induces the same orientation on the boundary as S. Thus, by Stokes’
theorem, ⁄⁄ ⁄⁄
(Ò ◊ F) · dS = (Ò ◊ F) · dSÕ .
S SÕ
To evaluate the integral on the right-hand-side, we first calculate the curl of F. We get:
1 2
Ò ◊ F = ≠x, 2x sin(x2 ), z ≠ x .
Next, we need to parametrize S Õ , which is the square with vertices (±1, ±1, ≠1). We
write – : D æ R3 with
and
–(u, v) = (u, v, ≠1).
The tangent vectors are
with the orientation induced by the parametrization. To evaluate the integral on the
right-hand-side, we calculate the exterior derivative:
which is simply the area of the region D. We then conclude, by Stokes’ theorem, that
the line integral of ÷ along ˆS is equal to the area of the region D.
5. Consider the two-form
1
Ê= 2 (x dy · dz + y dz · dx + z dx · dy),
(x + y + z 2 )3/2
2
Solution. (a) To show that it is closed, we calculate the exterior derivative. We get:
1 (x2 + y 2 + z 2 )3/2 ≠ 3x2 (x2 + y 2 + z 2 )1/2
dÊ =
(x2 + y 2 + z 2 )3
(x2 + y 2 + z 2 )3/2 ≠ 3y 2 (x2 + y 2 + z 2 )1/2
+
(x2 + y 2 + z 2 )3
(x2 + y 2 + z 2 )3/2 ≠ 3z 2 (x2 + y 2 + z 2 )1/2 2
+ dx · dy · dz
(x2 + y 2 + z 2 )3
1 1
2 2 2 2 2 2
2
= 2 3(x + y + z ) ≠ 3(x + y + z ) dx · dy · dz
(x + y 2 + z 2 )5/2
=0.
Therefore, Ê is closed.
(b) In fact, we already did such a surface integral in Exercise 5.6.4.5, but for complete-
ness we redo it here. We parametrize the sphere as always with spherical coordinates,
– : D æ R3 ,
D = {(◊, „) œ R2 | ◊ œ [0, fi], „ œ [0, 2fi]},
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 235
and
–(◊, „) = (sin(◊) cos(„), sin(◊) sin(„), cos(◊)).
The pullback of Ê is:
11
–ú Ê = sin(◊) cos(„)(cos(◊) sin(„)d◊ + sin(◊) cos(„)d„) · (≠ sin(◊)d◊)
1
+ sin(◊) sin(„)(≠ sin(◊)d◊)) · (cos(◊) cos(„)d◊) ≠ sin(◊) sin(„)d„)
2
+ cos(◊)(cos(◊) cos(„)d◊) ≠ sin(◊) sin(„)d„) · (cos(◊) sin(„)d◊ + sin(◊) cos(„)d„)
1 2
= sin3 (◊) cos2 („) + sin3 (◊) sin2 („) + sin(◊) cos2 (◊) d◊ · d„
= sin(◊)d◊ · d„.
Solution. The key here is recall some of the vector calculus identities that we encoun-
tered previously. From Lemma 4.4.9, we know that
Ò ◊ (f F) = (Òf ) ◊ F + f Ò ◊ F.
From the point of view of differential forms, this identity follows from the graded product
rule for the exterior derivative.
Using this identity, we can write:
⁄⁄ ⁄⁄ ⁄⁄
(f Ò ◊ F) · dS = ≠ (Òf ) ◊ F + Ò ◊ (f F).
– – –
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 236
The second term on the right-hand-side is the surface integral of the curl of the vector
field f F. Using Stokes’ theorem, we can rewrite it as a line integral:
⁄⁄ ⁄
Ò ◊ (f F) = f F · dr.
– ˆ–
Objectives
You should be able to:
• Determine and evaluate appropriate surface integrals and flux integrals in the context
of applications in science.
flv · n̂dS.
But... what is the area dS? We will come back to this in Section 7.2. Suppose that the tiny
region of D that is mapped to the tiny piece of surface by the parametrization is a rectangle,
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 237
with sides of lengths du and dv. We can think of these two sides as being vectors in the u and
v directions with length du and dv. We can think of these vectors as being mapped by the
parametrization to the rescaled tangent vectors duTu and dvTv . These two vectors span a
parallelogram in the tangent plane, with area given by |duTu ◊ dvTv | = |Tu ◊ Tv |dudv. The
idea is that the area of this parallelogram is a good approximation of the area dS of the tiny
piece of surface, since the tangent plane is a good approximation of the surface. (And, when
we sum over tiny pieces of surface and take the limit of the Riemann sum, this approximation
will become exact.) As a result, we can write
dS = |Tu ◊ Tv |dA
for the area of the tiny piece of surface, with dA = dudv. Therefore, we get that the rate of
flow of fluid through this tiny piece of surface in the normal direction is:
Tu ◊ Tv
flv · n̂|Tu ◊ Tv |dA =flv · |Tu ◊ Tv |dA
|Tu ◊ Tv |
=flv · ndA,
calculates what is known as the electric flux across the surface S. In fact, one of the
important laws in electromagnetism is Gauss’s law, which relates the electric charge to the
flux of an electric field. More precisely, if S is a closed surface in an electric field E, Gauss’s
law states that the net charge Q enclosed by the surface S is given by
⁄⁄
Q = ‘0 E · dS,
S
i.e. it is the electric flux through S rescaled by a constant ‘0 known as the “permittivity of
free space”. Here the surface S should be given the orientation of an outward normal vector.
Example 5.9.2 The electric flux and net charge of a point source. Let q be a point
charge at the origin. The electric force follows an inverse square law, that is, the magnitude of
the electric field produced by the charge is inversely proportional to the square of the distance
from the charge. More precisely, the electric field produced by the point charge is the vector
field
q
E(x, y, z) = (x, y, z).
4fi‘0 (x + y 2 + z 2 )3/2
2
Now suppose that you want to calculate the electric flux produced by the point charge
across a sphere of radius R centered at the origin. By Gauss’s law, this should calculate the
total charge enclosed by the sphere. Since we only have a point charge at the origin, with
charge q, we expect the electric flux to be given by q. Is that what we get?
The electric flux across the sphere in the outward direction is given by the flux integral
⁄⁄
E · dS.
S
We already calculated such surface integrals in Exercise 5.6.4.5 using spherical coordinates.
The result of this exercise applies here, with C = 4fi‘
q
0
. We thus conclude that
⁄⁄
q q
E · dS = 4fi = .
S 4fi‘0 ‘0
In particular, Gauss’s law states that the total charge enclosed by the sphere is
⁄⁄
q
Q = ‘0 E · dS = ‘0 = q.
S ‘0
Phew!
We note that the flux does not depend on the radius of the sphere, as it should. In fact,
we can go further. An argument very similar to Exercise 5.7.3.5 holds here as well, but using
the divergence theorem (which we will explore in Section 6.2) instead of Green’s theorem.
The conclusion of the argument is that the flux of the electric field of a point charge at the
origin across any closed surface that encloses the origin, not just the sphere, is always equal
to q, as it should by Gauss’s law. Cool! ⇤
The concept of flux is also used for instance in the study of heat flow. Suppose that the
temperature at a point (x, y, z) in a substance is given by the function T (x, y, z). Then the
heat flow is given the gradient of the temperature function, rescaled by a constant. More
precisely, the heat flow is
F = ≠KÒT,
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 239
where K is a constant called the conductivity of the substance. We are then often interested
in calculating the rate of heat flow across a surface S, which is given by the flux of the vector
field F across S: ⁄⁄ ⁄⁄
F · dS = ≠K ÒT · dS.
S S
5.9.3 Exercises
1. Consider a fluid moving with velocity v(x, y, z) = (≠y, x, 0) and constant mass density
fl(x, y, z) = fl0 with fl0 a positive constant. (As we saw in Exercise 2.1.3.2, this type of
fluid motion is a vortex in the (x, y)-plane.) Show that the rate of flow of fluid across
the cylinder x2 + y 2 = R2 , with ≠a Æ z Æ a, for some positive constants a, R, is zero.
Explain why this result makes sense, looking at the fluid motion.
Solution. As we saw in Exercise 2.1.3.2, this type of fluid motion is a vortex in
the (x, y)-plane. Below is, first, a sketch of the velocity vector field of the fluid in
3 dimensions. But since there is no motion in the z-direction, I also present a two-
dimensional plot of the vector field in the (x, y)-plane, which makes the vortex motion more
Figure 5.9.4 The vector field v(x, y) = (≠y, x), which corresponds to the projection of
the previous vector field in the (x, y)-plane.
Looking at the two-dimensional figure, we see that the motion is moving around in
a circular motion in the (x, y)-plane about the origin. In three dimensions, since there
is no motion in the z-direction, the fluid flow is circular about the z-axis. As such, if
we imagine a cylinder centered on the z-axis in the fluid, we see that there is no fluid
flowing across the surface of the cylinder. Thus we expect that the rate of flow of the
fluid across the cylinder should be zero. Let us show this.
To calculate the rate of flow across the cylinder S, we need to evaluate the surface
integral ⁄⁄
flv · dS.
S
and
–(u, ◊) = (R cos(◊), R sin(◊), u).
The tangent vectors are
It is pointing inward, but it does not matter anyway which orientation we choose as we
will show that the integral is zero. The surface integral is
⁄⁄ ⁄⁄
flv · dS = F(–(u, ◊)) · n dA
S ⁄⁄D
= (≠R sin(◊), R cos(◊), 0) · (≠R cos(◊), ≠R sin(◊), 0) dA
⁄⁄D
= (R2 sin(◊) cos(◊) ≠ R2 sin(◊) cos(◊)) dA
D
=0.
This is just as we expected: there is no rate of flow of the fluid across the cylinder.
2. Find the flux of the vector field
F(x, y, z) = (x, 0, z)
surface integral on both components, and add up the results to calculate the flux of the
vector field exiting the cone.
We parametrize S1 as –1 : D1 æ R3 with
with
–(r, ◊) = (r cos(◊), r sin(◊), 9 ≠ r2 ).
The tangent vectors are
3 4
r
Tr = cos(◊), sin(◊), ≠ Ô , T◊ = (≠r sin(◊), r cos(◊), 0).
9 ≠ r2
The normal vector is
A B
r2 r2
n = Tr ◊ T◊ = Ô cos(◊), Ô sin(◊), r .
9 ≠ r2 9 ≠ r2
It points upward in the z-direction, and thus outward of the cone, as we want to calculate
the flux exiting the cone. The surface integral is then
⁄⁄ ⁄⁄ A B
1 2 r2 r2
F · dS1 = r cos(◊), 0, 9 ≠ r2 · Ô cos(◊), Ô sin(◊), r dA
S1 D1 9 ≠ r2 9 ≠ r2
⁄⁄ A B
r3
= Ô cos2 (◊) + r 9 ≠ r2 dA
D1 9 ≠ r2
⁄ 3 ⁄ 2fi A B
r3
= Ô cos2 (◊) + r 9 ≠ r2 d◊dr
0 0 9 ≠ r2
⁄ 3A B
r3
=fi Ô + 2r 9 ≠ r2 dr
0 9 ≠ r2
⁄ 03 4
1u≠9 Ô
=fi Ô ≠ u du
9 2 u
⁄ 03 4
1 9
=fi ≠ u1/2 ≠ u≠1/2 du
9 2 2
5 6u=0
1 3/2 1/2
=fi ≠ u ≠ 9u
3 u=9
=36fi.
and
–(r, ◊) = (r sin(◊), r cos(◊), 0).
The tangent vectors are
and
–(u, ◊) = (cos(◊), sin(◊), u).
The tangent vectors are
n = Tu ◊ T◊ = (≠ cos(◊), ≠ sin(◊), 0)
This is pointing inward, so we change the sign of the normal vector. The surface integral
is then:
⁄⁄ ⁄⁄
E · dS = (cos(◊), 0, 0) · (cos(◊), sin(◊), 0) dA
– D
⁄ 2 ⁄ 2fi
= cos2 (◊) d◊du
≠2 0
CHAPTER 5. INTEGRATING TWO-FORMS: SURFACE INTEGRALS 244
⁄ 2
=fi du
≠2
=4fi.
We conclude, by Gauss’s law, that the net charge enclosed by the surface is
⁄⁄
Q = ‘0 E · dS = 4fi‘0 .
–
4. Suppose that the temperature distribution in a substance in R3 is given by the function
T (x, y, z) = x2 + xy.
Show that the rate of heat flow along any horizontal plane z = C, where C is a constant,
is zero. Explain why this is consistent with your expectation.
Solution. The heat flow is given by
To calculate the rate of heat flow across an horizontal plane z = C, we need to calculate
the surface integral of the heat flow F along the plane. In this process, we take the dot
product of F with the normal vector n to extract the normal component of F. Since the
plane is horizontal, its normal vector will point in the z-direction. But the vector field
F has a vanishing z-component; therefore, F · n = 0, and the surface integral along the
horizontal plane will vanish.
This is consistent with our expectation because the temperature distribution does
not depend on z. So it does not vary as we move in the z-direction; indeed, its gradient
has a vanishing z-component. As a result, there is no heat flowing through horizontal
planes, as we found.
Chapter 6
Objectives
You should be able to:
• Show that for integration of an exact one-form over a closed interval in R, it reduces to
the Fundamental Theorem of Calculus.
• Show that for integration of an exact one-form over a parametric curve in Rn , it reduces
to the Fundamental Theorem of line integrals.
• Show that for integration of an exact two-form over a closed bounded region in R2 , it
reduces to Green’s theorem.
• Show that for integration of an exact two-form over a parametric surface in R3 , it reduces
to Stokes’ theorem.
245
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 246
In mathematics, when we see something like this, we dig deeper and try to determine
whether it is a coincidence or not that all these integral theorems pretty much take the same
form. More often than not, such a coincidence is a hint that there is something going on
behind the scenes, that there is a unifying principle at play. This is precisely the case here.
The unifying principle is the mother of all integral theorems, known as the “generalized
Stokes’ theorem”. It states that the relationship above is very general. The precise statement
is the following.
Theorem 6.1.1 The generalized Stokes’ theorem. Let M be a k-dimensional oriented
manifold and ˆM its boundary (which is a (k ≠ 1)-dimensional manifold) with the induced
orientation. Let Ê be a (k ≠ 1)-form on M . Then
⁄ ⁄
dÊ = Ê.
M ˆM
The scope of this theorem is really awe-inspiring, at least in the eye of a mathematician.
We will not prove this theorem as it is beyond the scope of the class, but we can try to make
sense of it.
We know what a k-form is, at least over open subsets in Rn . The key object that we have
not defined and that appears in the statement of the theorem is the notion of a “manifold”,
which is fundamental in differential geometry. So let us say a few words about manifolds.
usually require that these transition functions, or coordinate changes, are differentiable (or
even smooth).
In the end, the key feature of a manifold is that locally, instead of doing calculations on
the space M itself, you can use the coordinate chart to do calculations on Rn instead. How
do we do this? We use the pullback! Indeed, our coordinate chart is an invertible map, so we
can pullback objects on M via the inverse of the coordinate chart to turn them into objects
on Rn , where we can do calculus. For instance, using pullback with respect to a coordinate
chart, we can define integration of an n-form on a region of an n-dimensional manifold via
integration of an n-form over a region in Rn , which is something that we studied in this class.
We use once again the fundamental principle of reducing something complicated to something
that we already know how to solve, and the pullback is there to help! How neat is this.
Most of the spaces that we encountered in this class, such as parametric curves, parametric
surfaces, etc. are examples of manifolds. But the definition of manifolds is much more general.
A key feature of manifolds is that they are defined “intrinsically”. When we talked about
parametric curves, or parametric surfaces, we introduced complicated geometry, but the way
we did it was by embedding a curve or a surface in a higher-dimensional space Rm . For
manifolds, you do not need to embed them into higher-dimensional spaces to get interesting
geometry; the geometry is intrinsic in the definition of a manifold.
But in the end, you already know many manifolds. Here are a few examples.
• The sphere (i.e. the surface of a ball) is a two-dimensional manifold, with no boundary.
• Parametric curves (the way we defined them) are one-dimensional manifolds, possibly
with boundary.
To end this section, we summarize in a table how the four integral theorems that we
already saw arise as special cases of the generalized Stokes’ theorem. We add a fifth integral
theorem to the table: the divergence theorem, which is the topic of the next section.
Table 6.1.2 Integral theorems as special cases of the generalized Stokes’ theorem
M Ê Integral theorem
Closed interval in R 0-form Fundamental theorem of calculus
Parametric curve in Rn 0-form Fundamental theorem of line integrals
Closed bounded region in R2 1-form Green’s theorem
Parametric surface in R3 1-form Stokes’ theorem
Closed bounded region in R3 2-form Divergence theorem
Next time someone tells you something about the fundamental theorem of calculus, you
can reply: “oh, I know this theorem, it’s just a special case of the generalized Stokes’ theorem”!
:-)
Objectives
You should be able to:
• For exact three-forms, rephrase the generalized Stokes’ theorem as the divergence
theorem in R3 .
• Summarize all the integral theorems of vector calculus as particular cases of the general-
ized Stokes’ theorem.
of E as being the orientation induced by the ambient vector space R3 . We write E+ for the
region with the canonical (right-handed twirl) orientation, and E≠ for the opposite (left-handed
twirl) orientation.
If E is canonically oriented, we define the induced orientation on the boundary ˆE
as corresponding to an outward pointing normal vector. ⌃
Example 6.2.2 Solid region bounded by a sphere in R3 . Let E µ R3 be the solid
region in R3 bounded by the sphere x2 + y 2 + z 2 = R2 . Its boundary ˆE is the sphere itself,
that is, the surface x2 + y 2 + z 2 = R2 . If we give E the canonical orientation given by a choice
of ordered basis on R3 corresponding to a right-handed twirl, then the induced orientation on
the sphere is that of a normal vector pointing outward. ⇤
With this definition of orientation, we can define the integral of a three-form over a region
in R3 .
Definition 6.2.3 Integral of a three-form over a closed bounded region in R3 . Let
E µ R3 be a solid region that consists of a closed surface and its interior. Let Ê = f dx·dy ·dz
be a three-form on an open subset U ™ R3 that contains E. If E has canonical orientation,
we define the integral of Ê over E as:
⁄ ⁄⁄⁄
Ê= f dV,
E E
where on the right-hand-side we mean the standard triple integral from calculus. If E has
opposite orientation, we define ⁄ ⁄⁄⁄
Ê=≠ f dV.
E E
Note that, as for the integral of two-forms over regions, the choice of basic three-form dx·dy·dz
(which is consistent with the ordering of the canonical orientation) when expressing Ê in
terms of the function f is important here, as integrals of three-forms are oriented, while triple
integrals are not. (See Remark 5.3.5.) ⌃
The definition reduces the evaluation of integrals of three-forms to triple integrals, which
you have encountered already in your previous calculus course. We will focus here on regions
that are recursively supported, as we did for regions in R2 (more general regions could be
expressed as unions of recursively supported regions). We say that:
1. A region E µ R3 is xy-supported (also called type 1) if there exists a region D in the
(x, y)-plane and two continuous functions z1 (x, y), z2 (x, y) such that
If the two-dimensional region D µ R2 is also supported on at least one of the two coordinates,
we say that the solid region E is recursively supported. In this case the triple integral can
be evaluated as an interated integral.
We note that in the case of rectangular regions, Fubini’s theorem still applies, and the
order of integration for the iterated integrals does not matter.
Example 6.2.4 Integral of a three-form over a recursively supported region. Let
E µ R3 be the solid region bounded by the planes x = 0, x = 2, y = 2, z = 0 and z = y. Find
the integral of the three-form
Ê = xyz dx · dy · dz
over E with canonical orientation.
By definition, we know that
⁄ ⁄⁄⁄
Ê= xyz dV.
E E
which shows that E is recursively supported. We can then write the triple integral as an
iterated integral, and evaluate:
⁄⁄⁄ ⁄ 2⁄ 2⁄ y
xyz dV = xyz dzdydx
E 0 0 0
⁄ 2⁄ 2 C Dz=y
z2
= xy dydx
0 0 2 z=0
⁄ 2⁄ 2
1
= xy 3 dydx
2 0 0
⁄ 2 C Dy=2
1 y4
= x dx
2 0 4 y=0
⁄ 2
=2 x dx
0
=4.
⇤
To conclude this section, we should show that our integration theory is invariant under
orientation-preserving reparametrizations, just as we did for two-forms in Subsection 5.3.2.
We will be brief here and simply state the result. Let E1 , E2 µ R3 be recursively supported
regions, and let „ : E2 æ E1 be a bijective and invertible function (that can be extended to a
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 251
where the integral on the right-hand-side is a surface integral of Ê over the boundary ˆE
realized as a parametric surface.
In vector calculus language, if F is the vector field associated to the two-form Ê, then
⁄⁄⁄ ⁄⁄
(Ò · F) dV = F · dS,
E ˆE
where the integral on the right-hand-side is the surface integral of the vector field F over the
boundary surface ˆE with normal vector pointing outward.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 252
We will not prove the divergence theorem here. We note that it follows directly from the
generalized Stokes’ theorem Theorem 6.1.1, just like our four other integral theorems. The
vector calculus translation follows directly from our dictionary between differential forms and
vector calculus concepts.
Remark 6.2.7 Just as for Green’s theorem, we can read the divergence theorem in two
different ways, starting from the left-hand-side or the right-hand-side. This results in two
potential applications: either to evaluate the volume integral of the divergence of a vector
field (or an exact three-form), or to evaluate the surface integral of a vector field (a two-form)
over a closed surface. However, as was the case for Green’s theorem, the divergence theorem
is mostly useful to evaluate surface integrals over closed surfaces by transforming them into
volume integrals over the interior of the region.
Example 6.2.8 Using the divergence theorem to evaluate the flux of a vector field
over a closed surface in R3 . Find the flux of the vector field
in the outward direction over the surface of the solid region E that lies above the (x, y)-plane
and below the surface z = 2 ≠ x ≠ y 3 , x œ [≠1, 1], y œ [≠1, 1].
We could try to evaluate the surface integral directly, but given how complicated the
vector field is, this would probably be a nightmare. Or we can use the divergence theorem,
which tells us that ⁄⁄ ⁄⁄⁄
F · dS = (Ò · F)dV.
ˆE E
The divergence of the vector field is
Ò · F = y + 1,
which is of course much simpler, so using the divergence theorem looks like a good strategy.
To evaluate the volume integral we need to write the solid region E as a recursively
supported region. Let us first look at what the solid region looks like. The surface Z = 2≠x≠y 3
is shown below.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 253
Figure 6.2.9 The surface z = 2 ≠ x ≠ y 3 over the rectangle [≠1, 1] ◊ [≠1, 1] in the (x, y)-plane.
The solid region E consists of the region bounded by the surface shown above, the four
sides of the box in the figure, and the bottom of the box. It can be written as an xy-supported
region:
E = {(x, y, z) œ R3 | x œ [≠1, 1], y œ [≠1, 1], 0 Æ z Æ 2 ≠ x ≠ y 3 }.
The volume integral can then be evaluated:
⁄⁄⁄ ⁄ 1 ⁄ 1 ⁄ 2≠x≠y3
(Ò · F)dV = (y + 1) dzdydx
E ≠1 ≠1 0
⁄ 1 ⁄ 1
3
= (y + 1) [z]z=2≠x≠y
z=0 dydx
≠1 ≠1
⁄ 1 ⁄ 1
= (2y ≠ xy ≠ y 4 + 2 ≠ x ≠ y 3 ) dydx
≠1 ≠1
⁄ 1 C Dy=1
2 xy 2 y 5 y4
= y ≠ ≠ + 2y ≠ xy ≠ dx
≠1 2 5 4 y=≠1
⁄ 1 3 4
18
= ≠ 2x dx
≠1 5
36
= .
5
Therefore, by the divergence theorem, the flux of F over the surface ˆE is equal to 36/5. ⇤
6.2.3 Exercises
1. Use the divergence theorem to find the surface integral of the two-form
Ê = 3xy 2 dy · dz + (ex z + yz 2 ) dz · dx + xy dx · dy
over the surface of the solid bounded by the cylinder y 2 + z 2 = 1 and the planes x = ≠1,
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 254
where the integral on the right-hand-side is over the solid region E with canonical
orientation. Thus instead of evaluating the surface integral, we can evaluate the volume
integral of the three-form dÊ over the solid region E.
We calculate the exterior derivative:
dÊ = (3y 2 + z 2 ) dx · dy · dz.
We note that the determinant of the Jacobian is r, which is positive, and hence the
integral is invariant under the change of coordinates (pullback). Thus we can rewrite the
integral as
⁄ ⁄
dÊ = „ú (dÊ)
E Õ
⁄E
= r3 (1 + 2 cos2 (◊)) du · dr · d◊.
EÕ
Therefore, the integral of Ê over the surface ˆE specified in the question is equal to fi.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 255
2. Use the divergence theorem to find the flux of the vector field
F(x, y, z) = (z + sin(y), y, ex )
Ò · F = 0 + 1 + 0 = 1.
which is simply the volume of the solid region E. As E is the interior of the sphere of
radius 4, we know right away that its volume is
4 256fi
fi(43 ) = ,
3 3
so we could conclude right away that the flux of the vector field across the sphere ˆE is
equal to 256fi
3 .
3. Use the divergence theorem to evaluate the surface integral of the two-form
over the surface of the solid bounded by the paraboloid z = 1≠x2 ≠y 2 and the (x, y)-plane,
with orientation given by an inward pointing normal vector.
Solution. First, we note that the problem is asking to evaluate the surface integral
with a normal vector pointing inward. Thus, if we denote the solid region by E and its
boundary surface by ˆE the divergence theorem tells us that
⁄ ⁄
Ê=≠ dÊ,
ˆE≠ E
where on the left-hand-side we mean the surface integral with normal vector point-
ing inward, while on the right-hand-side we mean the volume integral with canonical
orientation.
We calculate the exterior derivative:
We can either work in Cartesian or cylindrical coordinates here. But cylindrical coor-
dinates will make the calculation much easier (believe me: I first did it in Cartesian
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 256
We note that the determinant of the Jacobian is r, which is positive, and hence the
integral is invariant under the change of coordinates (pullback). Thus we can rewrite the
integral as
⁄ ⁄
dÊ = „ú (dÊ)
E Õ
⁄E
= 3r3 (2 + cos2 (◊)) dr · d◊ · du.
EÕ
⁄ 1 ⁄ Ô1≠x2
=3 Ô (3x2 + 2y 2 )(1 ≠ x2 ≠ y 2 ) dydx
≠1 ≠ 1≠x2
⁄ 1 ⁄ Ô1≠x2
=3 Ô (3x2 ≠ 3x4 + (≠5x2 + 2)y 2 ≠ 2y 4 ) dydx
≠1 ≠ 1≠x 2
Ô
⁄ 1 5 6
y= 1≠x 2
1 2
=3 y(3x2 ≠ 3x4 + (≠5x2 + 2)y 2 ≠ y 4 ) dx
≠1 3 5 Ô
y=≠ 1≠x2
⁄ 1 3 4
2 4 1 2 2 2 2 2
=6 1 ≠ x 3x ≠ 3x + (≠5x + 2)(1 ≠ x ) ≠ (1 ≠ x )
2 dx
≠1 3 5
⁄ 1 3 4
26 22 4
=6 1 ≠ x2 ≠ x4 + x2 + dx
≠1 15 15 15
⁄ fi/2 3 4
26 22 4
=6 cos2 (◊) ≠ sin4 (◊) + sin2 (◊) + d◊
≠fi/2 15 15 15
5fi
= .
4
Here I used the trigonometric substitution x = sin(◊), and to evaluate the last trigono-
metric integral one needs to use a whole bunch of trigonometric identities (or a computer
algebra system :-). We fortunately get the same result as before for the volume integral,
and we conclude as before that the surface integral is equal to minus this result, that is,
≠ 5fi
4 .
4. Show that the volume V of a solid region E µ R3 bounded by a closed surface ˆE can
be written as ⁄
V = x dy · dz,
ˆE
where the surface integral is evaluated with the orientation given by an outward pointing
normal vector.
Solution. We know that the volume of the region E is given by the integral of the
basic three-form Ê = dx · dy · dz over E with canonical orientation:
⁄
V = dx · dy · dz.
E
But Ê = dx·dy ·dz is exact, as it can be written as Ê = d÷ for the two-form ÷ = x dy ·dz.
Therefore, by the divergence the theorem, we know that
⁄ ⁄
dx · dy · dz = x dy · dz,
E ˆE
where the integral on the right-hand-side is the surface integral with orientation given by
an upward pointing normal vector.
5. We studied the two-form
1
Ê= 2 (x dy · dz + y dz · dx + z dx · dy)
(x + y 2 + z 2 )3/2
in Remark 4.6.6 and Exercise 5.8.3.5. Ê is defined on U = R3 \ {(0, 0, 0)}. We proved
in Exercise 5.8.3.5 that Ê is closed, but that it is not exact, by showing that its surface
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 258
integral along the sphere x2 + y 2 + z 2 = 1 is equal to 4fi. In this problem we show that
the surface integral of Ê is non-vanishing for all closed surface that contain the origin
(and always equal to 4fi), while it vanishes for all closed surfaces that do not pass through
or enclose the origin.
(a) Consider an arbitrary closed surface S0 with outward pointing normal vector that
does not contain or pass through the origin. Use the divergence theorem to show
that ⁄
Ê = 0.
S0
(b) Let S1 be an arbitrary closed surface with outward pointing normal vector that
contains the origin. Explain why the argument of (a) does not apply here.
(c) Let S1 be an arbitrary closed surface with outward pointing normal vector that
contains the origin. Let K be a sphere centered at the origin, with a radius small
enough that it is contained completely inside S1 . Give K the orientation of a
normal vector pointing outward (outward of the sphere K). Use the divergence
theorem to show that ⁄ ⁄
Ê= Ê.
S1 K
You have shown that the surface integral of Ê along any closed surface that contains
the origin is 4fi, while the surface integral of Ê along any closed surface that does not
enclose or pass through the origin is zero. Note that this is the argument that is needed
to show that, using Gauss’s law, the total charge contained within any closed surface
that encloses a point charge q at the origin is always equal to q -- see Example 5.9.2.
Solution. (a) We know that the two-form Ê is closed, that is, dÊ = 0. Furthermore,
Ê is defined on U = R3 \ {(0, 0, 0)}. If the surface S0 does not contain or pass through
the origin, then the surface S0 and the solid region E0 bounded by S0 lie within U , the
domain of definition of Ê. Therefore, by the divergence theorem,
⁄ ⁄
Ê= dÊ = 0.
S0 E0
(b) If S1 contains the origin in its interior, we cannot apply the divergence theorem
as in (a), since the solid region E1 bounded by S1 does not lie within U , the domain of
definition of Ê.
(c) We stated the divergence theorem only for solid regions that consisted of a single
closed surface and its interior, but it in fact applies to any closed bounded region. One
simply needs to make sure that each bounding surface is oriented with an outward
pointing normal vector, where “outward” means away from the solid region bounded by
the surfaces.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 259
In particular, we can consider the solid region E that is inside S1 , but outside the
sphere K. It is bounded by the two closed surfaces S1 and K. As the origin is inside
K, it is not within the solid region E. Therefore, E µ U , and the divergence theorem
applies to this solid region. In this case, the divergence theorem says that the sum of the
surface integral over the outer boundary S1 with outward pointing normal vector and
the inner boundary K with vector pointing away from the solid region (which means
that it is pointing inside the sphere K) is equal to the volume integral over E, which is
zero since dÊ = 0: ⁄ ⁄ ⁄
Ê+ Ê= dÊ = 0.
S1 ,out K,in E
We conclude that ⁄ ⁄
Ê=≠ Ê.
S1 ,out K,in
To get rid of the minus sign, we can reverse the orientation on K, and consider a normal
vector that points outward of the sphere K. We get:
⁄ ⁄
Ê= Ê,
S1 ,out K,out
Objectives
You should be able to:
• Use the Hodge star operator to rewrite the generalized Stokes’ theorem in Rn , which
can be rewritten as the divergence theorem in Rn .
Contrary to Green’s and Stokes’ theorem, the divergence theorem involves the divergence
of the vector field, not the curl. While the notion of curl of a vector field is not so easy to
generalize to Rn , the divergence can be generalized easily.
More precisely, let F(x1 , . . . , xn ) = (f1 , . . . , fn ) be a smooth vector field on U ™ Rn , with
f1 , . . . , fn : U æ R smooth functions. We can define the divergence of F naturally as
n
ÿ ˆfi ˆf1 ˆf2 ˆfn
Ò·F= = + + ... + .
i=1
ˆxi ˆx1 ˆx2 ˆxn
Now suppose that E µ Rn is a closed bounded region that consists of a closed (n ≠ 1)-
dimensional space ˆE and its interior. The integral of Ò · F over E is defined naturally in
calculus as a multiple (“n-tuple”) integral, which can be rewritten as an iterated integral if E
is recursively supported. The “surface” integral over ˆE can also be generalized; since ˆE is
a (n ≠ 1)-dimensional subspace in Rn , it has a well defined normal vector. If E is canonically
oriented (choose the canonical ordered basis on Rn ), we say that the induced orientation on
ˆE corresponds to an outward pointing normal vector, as for R3 . A natural question then
arise: does the divergence theorem generalize to any dimension? That is, is it true that
⁄ ⁄ ⁄ ⁄
··· (Ò · F) dVn = ··· (F · n)dVn≠1
E ˆE
¸ ˚˙ ˝ ¸ ˚˙ ˝
n times (n ≠ 1) times
where the integral on the left-hand-side is an n-tuple integral over the region E µ Rn , and
the right-hand-side is an integral of the vector field F over the parametrized surface ˆE with
normal vector pointing outward?
The answer is yes, and it again follows from the generalized Stokes’ theorem. But we need
to rewrite the generalized Stokes’ theorem a little bit to see this.
by definition of the Hodge star. So we can rewrite the generalized Stokes’ theorem for the
one-form ÷ as follows: ⁄ ⁄
d(ı÷) = ı÷.
E ˆE
This is really the same generalized Stokes’ theorem, but instead of writing it in terms of a
(n ≠ 1)-form Ê, we write it in terms of a one-form ÷.
Why would that be of any use? The advantage is that we can easily translate to vector
field concepts for all Rn , since we can establish a direct translation between one-forms and
vector fields regardless of the dimension.
where the hat notation means that we take the wedge product of all dxj ’s except the dxi .
Calculating the exterior derivative, we get:
n
ÿ ˆfi
d(ı÷) = (≠1)i≠1 ‰i · · · · · dxn
dxi · dx1 · · · · dx
i=1
ˆxi
n
ÿ ˆfi
= dx1 · · · · · dxn
i=1
ˆxi
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 262
⌅
As for
s
the right-hand-side of the generalized Stokes’ theorem, we need to rewrite the
integral ˆE ı÷ in terms of vector calculus objects. ˆE is a closed (n ≠ 1)-dimensional space
in Rn . We can think of it as a parametric space – : D æ Rn for some closed bounded region
D µ Rn≠1 , like we did for parametric curves in R2 and parametric surfaces in R3 . In this we
case, we can define the integral by pulling back using the parametrization. We claim that the
following lemma holds:
Lemma 6.3.2 Rewriting the right-hand-side. If the boundary space ˆE is realized as a
parametric space – : D æ Rn ,
⁄ ⁄ ⁄
ı÷ = ··· (F · n)dVn≠1 ,
ˆE
¸ ˚˙ D˝
(n ≠ 1) times
which is a multiple (“(n ≠ 1)-tuple”) integral over the closed bounded region D µ Rn≠1 . Here,
n is the normal vector to ˆE pointing outward.1
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 263
Proof. We will not prove this statement in general; we will only prove it for parametric curves
and surfaces. In fact, for parametric surfaces, this is basically the statement that was already
proven in Corollary 5.6.5; indeed, what we have in this case is a surface integral in R3 , and
because by Table 4.1.11 we know that the vector field associated to the two-form ı÷ is the
same as the vector field associated to the one-form ÷, the result of Corollary 5.6.5 still holds
here.
Let us then show that it holds for parametric curves in R2 . In this case, ÷ = f dx + g dy,
with associated vector field F = (f, g), and ı÷ = f dy ≠ g dx. Let – : [a, b] æ R2 be a
parametric curve representing the boundary curve ˆE. Thus we have:
⁄ ⁄ ⁄
ı÷ = ı÷ = –ú (ı÷).
ˆE – [a,b]
where both sides should be understood as multiple integrals over the corresponding regions.
To be precise, we need to specify what normal vector n we are using here. In R3 , we take
the normal vector n = Tu ◊ Tv induced by a parametrization of the surface ˆE (with the right
1
To be precise, we would need to specify what normal vector we are considering here, since it is not of unit
length. As we will see, in R3 the normal vector is the usual one n = Tu ◊ Tv induced by the parametrization
of the surface, while in R2 it is the normal vector that has the same norm as the tangent vector T to the
parametric curve.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 264
orientation); in R2 , we take the normal vector n = (y Õ (t), ≠xÕ (t)) in terms of a parmaetrization
–(t) = (x(t), y(t)) of the curve (with the right orientation), which has the same norm as the
tangent vector to the parametric curve, that is, |n| = |T| = (xÕ (t))2 + (y Õ (t))2 .
Remark 6.3.4 In some textbooks, the divergence theorem in R2 is simply called “another
form of Green’s theorem”. The reason is that it actually follows directly from Green’s theorem.
Recall that, given a vector field F = (f, g) in R2 , Green’s theorem states that
⁄⁄ ⁄
(Ò ◊ F) · e3 dA = (F · T) dt.
D ˆD
But if we rewrite this expression in terms of the original vector field F = (f, g), we get
⁄⁄ ⁄
(Ò · F) dA = (F · n) dt,
D ˆD
this physical interpretation of the curl and div precise by applying the Stokes’ and divergence
theorem (respectively) to the small sphere, and take a limit where the volume of the sphere
goes to zero. See for instance Section 4.4.1 in CLP 4 for this detailed calculation.
Objectives
You should be able to:
We consider the case where the vector field F is the heat flow. Recall the context from
Subsection 5.9.2. Suppose that the temperature at a point (x, y, z) in an object (or substance)
is given by the function T (x, y, z). The heat flow is given the gradient of the temperature
function, rescaled by a constant K known as the conductivity of the substance:
F = ≠KÒT.
Now consider any closed surface ˆE, with the surface ˆE and its interior within the object.
The amount of heat flowing across the surface ˆE is given by the flux of the vector field F
across ˆE: ⁄⁄ ⁄⁄
(F · n)dA = ≠K (ÒT · n) dA.
ˆE ˆE
If we are interested in the amount of heat entering the solid region E (instead of flowing
across the surface in the outward direction), we change the sign of the flux integral. Then, by
the divergence theorem, the amount of heat entering the solid region E can be rewritten as a
volume integral:
⁄⁄ ⁄⁄⁄
K (ÒT · n) dA =K (Ò · ÒT ) dV
ˆE ⁄⁄⁄E
=K Ò2 T dV,
E
“spectator variable” here; we still consider the operator Ò2 in R3 , in terms of the variables
(x, y, z). So we can go through all the steps above, and we obtain that the amount of heat
entering the solid region E during a small (infinitesimal) amount of time dt is
⁄⁄⁄
Kdt Ò2 T dV.
E
2 2 2
Here the Laplacian is only in the variables (x, y, z), that is, Ò2 = ˆx ˆ
2 + ˆy 2 + ˆz 2 .
ˆ ˆ
To proceed further, we need a little bit of physics. It is known in physics that the amount
of heat energy required to raise the temperature of an object by T is given by CM T ,
where M is the mass of the object and C is constant known as the “specific heat” of the
material. Now consider the object consisting of the solid region E. In a small (infinitesimal)
amount of time dt, the temperature changes by ˆT (x,y,z,t)
ˆt dt. If we consider an infinitesimal
volume element dV in E, and fl(x, y, z) is the mass density of the solid region, then the mass
of the volume element is fldV . Thus the heat energy required to change the temperature of
the object in the time interval dt is
ˆT
Cfl dV dt.
ˆt
We then sum over all volume elements, i.e. integrate over E, to get that the total heat energy
required to change the temperature during the time interval dt is
⁄⁄⁄
ˆT
Cdt fl dV.
E ˆt
Assuming that the object is not creating heat energy itself, this heat energy should be
equal to the amount of heat entering the solid region E through the boundary surface ˆE
during the time interval dt, which is what we calculated previously. We thus obtain the
equality: ⁄⁄⁄ ⁄⁄⁄
ˆT
Cdt fl dV = Kdt Ò2 T dV.
E ˆt E
We can cancel the time interval dt on both sides. Rewriting both terms on the same side of
the equality, we get: ⁄⁄⁄ 3 4
2 ˆT
KÒ T ≠ Cfl dV = 0.
E ˆt
But this must be true for all solid regions E within the object, and for all times t. From this
we can conclude that the integrand must be identically zero:
ˆT
KÒ2 T = Cfl .
ˆt
This equation is generally rewritten as
ˆT (x, y, z, t)
= –Ò2 T (x, y, z, t),
ˆt
2 2 2
where – = Cfl
K
is called the “thermal diffusivity”, and Ò2 = ˆx
ˆ
2 + ˆy 2 + ˆz 2 .
ˆ ˆ
This equation is very famous: it is known as the heat equation. As mentioned in the
Wikipedia page on “Heat equation”,
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 267
The importance of the heat equation goes beyond physics and heat flow. It has a wide range
of applications, from the physics of heat flow of course, to probability theory, to financial
mathematics, to quantum mechanics, to image analysis in computer science. A generalization
of the heat equation is also behind the famous proof of the Poincare conjecture by Pereleman
in 2003 (the only Millenium Prize Problem that has been solved so far). I encourage you to
have a look at the wikipedia page on the heat equation!
6.4.2 The divergence theorem in Rn and Green’s first and second identities
We now consider the divergence theorem in Rn . Let F be a vector field, ˆE a closed (n ≠ 1)-
dimensional subspace with normal vector pointing outward, and E the region of Rn consisting
of ˆE and its interior with canonical orientation. The divergence theorem in Rn is the
statement that ⁄ ⁄ ⁄ ⁄
· · · (Ò · F)dVn = ··· (F · n)dVn≠1 .
¸ ˚˙ E˝ ¸ ˚˙ ˆE˝
n times (n ≠ 1) times
Using this theorem, we can prove the following two identities, known as Green’s first and
second identities.
Lemma 6.4.1 Green’s first identity. Let f, g : Rn æ R be functions with continuous
partial derivatives. Let E and ˆE be as above. Then
⁄ ⁄ ⁄ ⁄ ⁄ ⁄
2
··· f Ò g dVn = ··· f Òg · n dVn≠1 ≠ ··· (Òf · Òg) dVn .
¸ ˚˙ E˝ ¸ ˚˙ ˆE˝ ¸ ˚˙ E˝
n times (n ≠ 1) times n times
Proof. We consider the divergence theorem in Rn with vector field F = f Òg. By the third
identity in Lemma 4.4.9, we know that
Ò · (f F) = (Òf ) · F + f Ò · F.
Thus
Ò · (f Òg) =Òf · Òg + f Ò · Òg
=Òf · Òg + f Ò2 g.
This may not be obvious at first, but Green’s first identity is essentially the equivalent of
integration by parts in higher dimension. Basically, integration by parts can be written as
⁄ b -b ⁄ b
-
f dg = f g - ≠ g df.
a a a
Green’s first identity generalizes this statement for the n-tuple integral of the function f Ò2 g
over a closed bounded region E µ Rn .
Lemma 6.4.2 Green’s second identity. Let f, g : Rn æ R be functions with continuous
partial derivatives. Let E and ˆE be as above. Then
⁄ ⁄ ⁄ ⁄
2 2
··· (f Ò g ≠ gÒ f ) dVn = ··· (f Òg ≠ gÒf ) · n dVn≠1 .
¸ ˚˙ E˝ ¸ ˚˙ ˆE˝
n times (n ≠ 1) times
Proof. Green’s second identity follows from the first identity. Using the first identity, we know
that
⁄ ⁄ ⁄ ⁄
2 2
··· (f Ò g ≠ gÒ f ) dVn = ··· (f Òg ≠ gÒf ) · n dVn≠1
¸ ˚˙ E˝ ¸ ˚˙ ˆE˝
n times (n ≠ 1) times
⁄ ⁄
≠ ··· (Òf · Òg ≠ Òg · Òf ) dVn .
¸ ˚˙ E˝
n times
But Òf · Òg = Òg · Òf , and hence the last term vanishes. The result is Green’s second
identity. ⌅
Green’s identities are quite useful in mathematics. There is in fact also a third Green’s
identity, but it is beyond the scope of this class. Have a look at the Wikipedia page on
“Green’s identities” if you are interested!
6.4.3 Exercises
1. Recall that a function g : U æ R with U ™ Rn is harmonic on U if it is a solution to
q ˆ2
the Laplace equation, that is, Ò2 g = 0 on U , where Ò2 = ni=1 ˆx 2 . Use Green’s first
i
identity to show that if g is harmonic on U , and E µ U (with E and ˆE as usual), then
⁄ ⁄
··· Òg · n dVn≠1 = 0.
ˆE
¸ ˚˙ ˝
(n ≠ 1) times
But Ò(1) = 0, since the gradient of a constant function necessarily vanishes. Furthermore,
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 269
Solution. We consider Green’s first identity again, but now with f = g. It reads:
⁄ ⁄ ⁄ ⁄ ⁄ ⁄
··· gÒ2 g dVn = ··· gÒg · n dVn≠1 ≠ ··· (Òg · Òg) dVn .
¸ ˚˙ E˝ ¸ ˚˙ ˆE˝ ¸ ˚˙ E˝
n times (n ≠ 1) times n times
vanishes, since the integrand is identically zero on the surface ˆE over which we are
integrating. As a result, Green’s first identity becomes
⁄ ⁄
··· (Òg · Òg) dVn = 0.
¸ ˚˙ E˝
n times
Objectives
You should be able to:
• Determine which integral theorem may be useful to evaluate certain type of line and
surface integrals.
CHAPTER 6. BEYOND ONE- AND TWO-FORMS 270
We have seen five integral theorems so far, all particular cases of the generalized Stokes’
theorem:
3. Green’s theorem;
4. Stokes’ theorem;
One of the main difficulties with the integral theorems of calculus is to determine which
theorem may be helpful in a given situation. In this section I list a few typical situations
for which integral theorems may be useful, highlighting the main applications of the integral
theorems. You can use this list as a rule of thumb.
Strategy 6.5.1 Integral theorems: when to use what.
• You want to evaluate a line integral along a curve in Rn for an exact one-form (or the
gradient of a function): Fundamental Theorem of line integrals (Section 3.4).
• You want to evaluate a line integral along a closed curve in R2 : Green’s theorem
(Section 5.7).
• You want to evaluate a line integral along a closed curve in R3 : Stokes’ theorem
(Section 5.8).
• You want to evaluate a surface integral along a surface in R3 for an exact two-form (or
the curl of a vector field): Stokes’ theorem (Section 5.8).
Objectives
You should be able to:
• Determine the arc length of a parametrized curve in Rn using an unoriented line integral.
271
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 272
From the point of view of vector fields, the orientation of the integral is encapsulated in the
choice of tangent vector T. However, the way it is formulated, the tangent vector includes
more information than just the orientation, as it also has a non-trivial norm specified by the
parametrization. To isolate the oriented nature of the integral, we normalize the tangent
vector, and define the unit tangent vector
T Ò
T̂ = , |T| = (xÕ1 (t))2 + . . . + (xÕn (t))2 .
|T|
With this formulation, we see that the choice of orientation is completely encapsulated in the
expression F(–(t)) · T̂(t), which is function of t which depends on the choice of direction on
the parametric curve.
We can now see how unoriented line integrals can be naturally defined: we simply replace
the function F(–(t)) · T̂(t), constructed out of a vector field and a choice of orientation on the
curve, by an arbitrary function f (–(t)) that does not depend on a choice of orientation. This
leads to the following definition.
Definition 7.1.1 Unoriented line integrals. Let – : [a, b] æ Rn be a parametric curve,
with image curve C = –([a, b]) µ Rn , and let f : C æ R be a continuous function. We define
the unoriented line integral of f along the curve C to be
⁄ ⁄ b ⁄ b Ò
f ds = f (–(t))|T(t)| dt = f (–(t)) (xÕ1 (t))2 + . . . + (xÕn (t))2 dt.
C a a
⌃
A similar calculation as in the proof of Lemma 3.3.5 shows that unoriented line integrals
are invariant under reparametrizations, regardless of whether the reparametrization preserves
the orientation or not (the integrals are unoriented). What this means is that the line integral
does not depend on the choice of parametrization, but only on the image curve C. This is
why we wrote ⁄
f ds
C
for the unoriented line integral of the function f along the curve C µ Rn , without specifying
the parametrization –, since the integral is independent of the choice of parametrization.
Remark 7.1.2 We note that even though the notation “ds” is similar to the notation we
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 273
used for one-forms, the line element is not a one-form. For instance,
⁄ ⁄
ds = ds,
C+ C≠
i.e. the integral remains the same if we change the orientation of the curve, which would not
be the case if ds was a one-form.
Example 7.1.3 An example of an unoriented line integral. Evaluate the unoriented
line integral ⁄
xy 6 ds,
C
As ◊ œ [≠fi/2, fi/2], we are parametrizing the right half of the circle, as required. We do not
need to check here whether the parametrization induces the right orientation on the curve, as
we do not care about the orientation whatsoever: the integral is unoriented.
To evaluate the unoriented line integral, we need the line element ds. We calculate:
Ò
ds = (xÕ (◊))2 + (y Õ (◊))2
Ò
= 4 sin2 (◊) + 4 cos2 (◊)
=2.
image curve C = –([a, b]) µ Rn . The arc length of C is given by the unoriented line integral
⁄ ⁄ bÒ
ds = (xÕ1 (t))2 + . . . + (xÕn (t))2 dt.
C a
⌃
You may have seen this formula before for the arc length, at least in or R2
It is R3 .
straightforward to justify that it calculates the arc length of the curve, using the standard
slicing argument from integral calculus. Consider a small curve segment between two points
P (–(t)) and Q(–(t + dt)). The length ds of this curve segment can be approximated by the
length of the line joining the two points, which can be written as
- -
ds ƒ |–(t + dt) ≠ –(t)| ƒ -–Õ (t)- dt,
where on the right-hand-side we kept only terms of first-order in dt. As –Õ (t) = T(t), we
recover the formula above for the line element. Finally, we sum over line elements and take
the limit of an infinite number of line element of infinitesimal size, which turns the sum into
the definite integral
⁄ b ⁄ bÒ
|T(t)| dt = (xÕ1 (t))2 + . . . + (xÕn (t))2 dt.
a a
Note that this also justifies our definition of unoriented line integrals in general above; it is
constructed via the same slicing process, but where we also introduce a function f evaluated
the point –(t) where we calculate the line element. We then sum over slices and take the limit
of infinite number of infinitesimal slices as usual, and the integral of the function over the
parametric curve becomes Definition 7.1.1.
Example 7.1.5 Calculating the arc length of a parametric curve. Find the length of
the parametric curve – : [0, 2] æ R3 with
–(t) = (1, t2 , t3 ).
We first calculate the line element ds:
Ò
ds = (xÕ (t))2 + (y Õ (t))2 + (z Õ (t))2 dt
= 4t2 + 9t4 dt
=t 4 + 9t2 dt,
Ô
where in the last line we used t2 = t since we know that t œ [0, 2] and hence it is positive.
The arc length is thus given by
⁄ ⁄ 2
ds = t 4 + 9t2 dt
C 0
⁄
1 40 Ô
= u du
18 4
1
= (403/2 ≠ 43/2 )
27
8
= (103/2 ≠ 1),
27
2
where we used the substitution u = 4 + 9t , du = 18t dt. ⇤
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 275
7.1.3 Exercises
1. Find the arc length of the circular helix – : [0, 3] æ R3 with
Solution. To find the arc length, we first calculate the line element ds. The tangent
vector is
T(t) = (1, ≠2 sin(t), 2 cos(t)).
Its norm is Ò Ô
|T(t)| = 1 + 4 sin2 (t) + 4 cos( t) = 5.
Thus the line element is Ô
ds = 5 dt.
We then calculate the arc length:
⁄ ⁄ 3Ô
ds = 5 dt
– 0
Ô
=3 5.
2. Show that the arc length of the curve C at the intersection of the surfaces x2 = 2y
and 3z = xy between the origin and the point (6, 18, 36) is the answer to the ultimate
question of life, the universe, and everything.
Solution. We first parametrize the curve as – : [0, 6] æ R3 with
A B
t2 t 3
–(t) = t, , .
2 6
=42,
which is of course the answer to the ultimate question of life, the univers, and everything!
:-)
3. Evaluate the unoriented line integral
⁄
(xz + e≠y ) ds,
C
where C is the line segment between the origin and the point (1, 2, 3).
Solution. We parametrize C as – : [0, 1] æ R3 with
Show that this is consistent with our definition of arc length in this section.
Solution. From our point of view, we realize the curve as the parametric curve – :
[a, b] æ R2 with
–(t) = (t, f (t)).
Then the tangent vector is
T(t) = (1, f Õ (t)),
with norm Ò
|T(t)| = 1 + (f Õ (t))2 .
So our arc length formula is
⁄ ⁄ b ⁄ bÒ
ds = |T(t)| dt = 1 + (f Õ (t))2 dt,
C a a
Objectives
You should be able to:
dS = |Tu ◊ Tv | dA.
With this rewriting, we can think of the surface integral as integrating the expression
(F(–(u, v)) · n̂), which is a function of the parameters (u, v) that depends on the choice
of orientation via the unit normal vector n̂.
Unoriented surface integrals are then naturally defined by replacing the orientation-
dependent function (F(–(u, v)) · n̂) by an arbitrary function f (–(u, v)) that does not depend
on a choice of orientation on the surface. We obtain the following definition of unoriented
surface integrals.
Definition 7.2.1 Unoriented surface integrals. Let – : D æ R3 be a parametric surface,
with –(u, v) = (x(u, v), y(u, v), z(u, v)), and image surface S = –(D) µ R3 . The tangent
vectors are
ˆ– ˆ–
Tu = , Tv = .
ˆu ˆv
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 279
⌃
The justification for this definition of the surface area comes from the standard slicing
process of integral calculus, as usual. Consider a small rectangle at position (u, v) within the
domain D, and with sides of length du and dv. The area of this small rectangle is dA = dudv.
The parametrization – maps the small rectangle to a small region dS within the surface S.
This region is now curved, and not necessarily rectangular. However, one can approximate its
area by replacing it by the parallelogram in the tangent plane at the point –(u, v) spanned by
the vectors duTu and dvTv . The area of this parallelogram is dS = |Tu ◊ Tv |dudv. Finally,
we sum over all small regions within the domain D, and take a limit of an infinite number
of infinitesimal regions, which turns the approximate calculation of the surface area into an
exact one. The result of this limit process is the double integral
⁄⁄
|Tu ◊ Tv | dA.
D
Admittedly, the justification is a little bit hand-wavy here, but it can be made precise.
This also justifies our general construction of unoriented surface integrals in Definition 7.2.1:
all that we add to the slicing process is a function on the surface S that we evaluate at each
point –(u, v) on S before summing over slices to turn the calculation into a double integral.
Example 7.2.4 Calculating the surface area of a parametric surface. Find the
surface area of the part of the plane x + 2y + z = 1 that lies within the cylinder x2 + y 2 = 4.
As we know that the region will be within the cylinder x2 +y 2 = 4, we use polar coordinates
to parametrize the surface. We write down the parametrization – : D æ R3 with
and
–(r, ◊) = (r cos(◊), r sin(◊), 1 ≠ r cos(◊) ≠ 2r sin(◊)).
We need to find the surface element dS. The tangent vectors are
7.2.3 Exercises
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 282
1. Find the surface area of the part of the surface z = 4 ≠ 2x2 + y over the triangle with
vertices (0, 0), (2, 0), (2, 2) in the (x, y)-plane.
Solution. First, we notice that the triangle in the (x, y)-plane can be realized as the
x-supported region x œ [0, 2], 0 Æ y Æ x. We can then parametrize the surface as
– : D æ R3 with
D = {(u, v) œ R2 | u œ [0, 2], 0 Æ v Æ u}
and
–(u, v) = (u, v, 4 ≠ 2u2 + v).
The tangent vectors are
and 3 4
rH
–(r, ◊) = r cos(◊), r sin(◊), .
R
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 283
since r œ [0, R] and hence is positive. We then calculate the surface area:
Û
⁄⁄ ⁄ R ⁄ 2fi
H2
dS = +1 r d◊dr
S R2 0 0
Û
⁄ R
H2
=2fi +1 r dr
R2 0
Û
H2
=fiR2 +1
R2
=fiR H 2 + R2 .
3. Evaluate the unoriented surface integral
⁄⁄
xy dS,
S
where S is the part of the plane x + 2y + z = 4 that lies in the first octant.
Solution. We first need to parametrize the surface. We want the part of the plane that
lies in the first octant, so we must have x Ø 0, y Ø 0, and z Ø 0. The equation of the
plane is z = 4 ≠ x ≠ 2y. Since z Ø 0 and y Ø 0, we know that x cannot be more than 4.
So we take x œ [0, 4]. Then, since z Ø 0, we know that 4 ≠ x ≠ 2y Ø 0, which means that
2y Æ 4 ≠ x, that is, y Æ 2 ≠ x2 . So we know that 0 Æ y Æ 2 ≠ x2 .
Now that we identified the region in the (x, y)-plane over which the part of the plane
is, we can parametrize it as – : D æ R3 , with
u
D = {(u, v) œ R2 |u œ [0, 4], 0 Æ v Æ 2 ≠ },
2
and
–(u, v) = (u, v, 4 ≠ u ≠ 2v) .
The tangent vectors are
whose norm is Ô
|Tu ◊ Tv | = 1 + 22 + 1 = 6.
We calculate the unoriented surface integral:
⁄⁄ Ô ⁄ 4 ⁄ 2≠ u2
xy dS = 6 uv dvdu
S
Ô ⁄0 0 3 4
6 4 u 2
= u 2≠ du
2 0 2
Ô ⁄ A B
6 4 2 u3
= 4u ≠ 2u + du
2 0 4
Ô A 4
B
6 2 4
= 2(42 ) ≠ 43 +
2 3 16
Ô
8 6
= .
3
4. Consider the surface S = {y = f (x)} µ R3 , where f : R æ R is a smooth function, and
x œ [0, 2], z œ [0, 1]. Show that the surface area of S is equal to the arc length of the
curve y = f (x), x œ [0, 2], in the (x, y)-plane.
Solution. We can parametrize the surface S by – : D æ R3 with
and
–(u, v) = (u, f (u), v).
The tangent vectors are
But this is precisely the formula for the arc length of the parametric curve — : [0, 2] æ R2
with —(u) = (u, f (u)), which is the curve {y = f (x)} µ R2 with x œ [0, 2].
By the way, this result is not surprising. Indeed, the surface S is basically just the
curve y = f (x) extended uniformly in the z-direction. So the surface area should be
the arc length of the curve times the length of the surface in the z-direction. Since S
is defined with z œ [0, 1], the length in the z-direction is just 1, so we the surface area
should be equal to the arc length of the curve, as we obtained.
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 285
5. In single-variable calculus, you saw that the surface area of the solid of revolution obtained
by rotating the curve y = f (x), x œ [a, b] (and a Ø 0), about the y-axis is given by the
definite integral
⁄ b Ò
A = 2fi x 1 + (f Õ (x))2 dx.
a
Show that this is consistent with our formula for the surface area of parametric surfaces.
Solution. The surface S obtained by rotating the curve y = f (x), x œ [a, b], about the
y-axis, can be parametrized as follows. First, we can parametrize the curve y = f (x) in
the (x, y)-plane as (u, f (u), 0), u œ [a, b]. When we rotate the curve about the y-axis,
for a fixed value of u, the point (u, 0) in the (x, z)-plane gets rotated about the y-axis
around a circle of radius u. So we should replace (u, 0) by (u cos(◊), u sin(◊)). This gives
a parametrization for the surface S as – : D æ R3 with
with
–(u, ◊) = (u cos(◊), f (u), u sin(◊)).
The tangent vectors are
Its norm is
Ò Ò
|Tu ◊ T◊ | = u2 (f Õ (u))2 cos2 (◊) + u2 (f Õ (u))2 sin2 (◊) + u2 = u 1 + (f Õ (u))2 ,
since u œ [a, b], a Ø 0, and hence u is positive. The surface area is then:
⁄⁄ ⁄ b ⁄ 2fi Ò
dS = u 1 + (f Õ (u))2 d◊du
S a 0
⁄ b Ò
=2fi u 1 + (f Õ (u))2 du,
a
Objectives
You should be able to:
• Determine and evaluate unoriented line and surface integrals in the context of applications
in science.
Now what if we want to calculate the centre of mass of a wire that is bent and twisted in
Rn ? We apply the same idea, but thinking of the wire as a parametric curve – : [a, b] æ Rn ,
with –(t) = (x1 (t), . . . , xn (t)) and image curve C = –([a, b]). Let fl : C æ R be the mass
density of the wire, which we assume to be continuous. We slice the wire (the image curve C)
into small curve segments. Let ds (the line element from Subsection 7.1.1) be the length of
a typical curve segment located at –(t). Its mass is dm = fl(–(t)) ds. By summing over all
curve segments and taking the limit of an infinite number of segments of infinitesimal length,
we calculate the total mass of the wire as an unoriented line integral:
⁄ ⁄ b
m= fl ds = fl(–(t))|T(t)| dt.
C a
To get the centre of mass, we also need to calculate the first moments of the wire. Here, we
are working in Rn . There are n first moments, one in each coordinate x1 , . . . , xn . The first
moments of the curve segment are given by
xk (t)fl(–(t)) ds, k = 1, . . . , n.
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 287
Summing over curve segments, and taking the limit as usual, we obtain that the position of
the centre of mass of the wire is given by the point in Rn with coordinates x̄1 , . . . , x̄n , with:
⁄ ⁄
1 1 b
x̄k = xk fl ds = xk (t)fl(–(t))|T(t)| dt, k = 1, . . . , n.
m C m a
In other words, to find the centre of mass of the wire, we need to evaluate the unoriented line
integrals corresponding to the total mass of the wire and its n first moments.
Note that the centre of mass of the wire will not generally be located on the wire itself,
since the wire is twisted and bent in the ambient space Rn ; this will be clear in the next
example.
Example 7.3.1 Finding the centre of mass of a wire in R2 . A wire is bent into the
semi-circle x2 + y 2 = 4, y Ø 0. Its mass density is given by the function
fl(x, y) = 4 ≠ y
=0,
To make sure that this is consistent with our expectation, we can find a numerical value for
the location of the centre of mass. We get:
Example 7.3.2 Finding the centre of mass of a sheet in R3 . Suppose the a sheet of
paper is bent in the shape of the cylinder x2 + y 2 = 9, with z œ [0, 1]. Suppose that its mass
density is given by the function
fl(x, y, z) = z + 1.
Find the centre of mass of the cylinder.
First, let us see what we expect. The cylinder is heavier at the top than at the bottom.
However, its mass density has circular symmetry above the z-axis, as it only depends on
the height z on the cylinder. Thus we expect the centre of mass to be in the middle of the
cylinder, i.e. with coordinates (0, 0, z̄). Furthermore, we expect its z-coordinate to be a little
bit higher than half-way up the cylinder, since the cylinder is heavier at the top than at the
bottom. So we expect 0.5 < z̄ < 1.
To calculate the required unoriented surface integrals, we first parametrize the surface as
– : D æ R3 with
D = {(u, ◊) œ R2 | u œ [0, 1], ◊ œ [0, 2fi]}
and
–(u, ◊) = (3 cos(◊), 3 sin(◊), u).
The tangent vectors are
Its norm is Ò
|Tu ◊ T◊ | = 9 cos2 (◊) + 9 sin2 (◊) = 3,
and thus the surface element is
dS = 3dud◊.
We calculate the total mass of the sheet of paper. We get:
⁄⁄
m= fl dS
S
⁄ 1 ⁄ 2fi
= (u + 1)3 d◊du
0 0
⁄ 1
=6fi (u + 1) du
0
=9fi.
Since 59 ƒ 0.556,, this is consistent with our expectation that the centre of mass should be on
the z-axis, a little bit higher than half-way up the cylinder, which is at z = 0.5. ⇤
7.3.3 Exercises
1. Find the centre of mass of a wire that is bent in the shape of a circle of radius R centered
at the origin in R2 , with mass density fl(x, y) = 2R ≠ y. Is your result consistent with
your expectations?
Solution. Let us first see what we expect. As the wire is bent in a circle centered at
the origin, if the mass density was constant (or symmetric about the x- and y-axes), the
centre of mass would be at the origin. However, the mass density is not constant here: it
decreases linearly with y. It is constant in the x-direction, so we expect that centre of
mass to lie on the y-axis. As the mass decreases as y increases, we expect the centre of
mass to be at a position (0, ȳ), with ≠R < ȳ < 0.
We parametrize the circle of radius R as – : [0, 2fi] æ R2 with –(◊) = (R cos(◊), R sin(◊)).
The tangent vector is
T(◊) = (≠R sin(◊), R cos(◊)).
Its norm is Ò
|T(◊)| = R2 sin2 (◊) + R2 cos2 (◊) = R.
The total mass of the wire is
⁄ ⁄ 2fi
m= fl ds = (2R ≠ R sin(◊))Rd◊
C 0
CHAPTER 7. UNORIENTED LINE AND SURFACE INTEGRALS 291
=4fiR2 .
Find the moments of inertia of a wire that lies along the line 2x + y = 5 between x = 0
and x = 1, with density fl(x, y) = x.
Solution. We parametrize the line as – : [0, 1] æ R2 with
Ô ⁄ 1
= 5 (25t ≠ 20t2 + 4t3 ) dt
0
Ô 3 25 20 4
= 5 ≠ +1
2 3
Ô
41 5
= .
6
Second,
⁄
Iy = x2 fl ds
C
Ô ⁄ 1 3
= 5 t dt
0
Ô
5
= .
4
3. Consider a sheet of aluminium foil that is bent in the shape of a paraboloid z = x2 + y 2
with z œ [0, 1]. Suppose that it mass density is fl = k where k is a constant. Find the
total mass of the sheet and its centre of mass. Does it agree with your expectation?
Solution. As the mass density is constant, and the paraboloid has circular symmetry
about the z-axis, we expect the centre of mass to lie on the z-axis, somewhere between
z = 0 and z = 1.
We parametrize the paraboloid as – : D æ R3 with
and
–(r, ◊) = (r cos(◊), r sin(◊), r2 ).
The tangent vectors are
Its norm is
Ò
|Tr ◊ T◊ | = 4r4 cos2 (◊) + 4r4 sin2 (◊) + r2 = r 4r2 + 1.
as expected. Next,
⁄⁄
1
ȳ = yfl dS
m S
⁄ 1 ⁄ 2fi
k
= r2 sin(◊) 4r2 + 1 d◊r
m 0 0
=0,
as expected. Finally,
⁄⁄
1
z̄ = zfl dS
m S
⁄ 1 ⁄ 2fi
6k
= ! " r3 4r2 + 1 d◊r
fik 53/2 ≠ 1 0 0
⁄ 1
12
= ! 3/2 " r3 4r2 + 1 dr
5 ≠1 0
⁄ 5
3 Ô
= ! 3/2 " (u ≠ 1) u du
8 5 ≠1 1
3 4
3 2 5/2 2 2 3/2 2
= ! " 5 ≠ ≠ 5 +
8 53/2 ≠ 1 5 5 3 3
1 55/2 + 1
= .
10 53/2 ≠ 1
List of results
294
APPENDIX A. LIST OF RESULTS 295
Chapter 4 k-forms
Lemma 4.1.6 Antisymmetry of basic k-forms
Lemma 4.2.6 Comparing Ê · ÷ to ÷ · Ê
Lemma 4.2.7 The wedge product of two one-forms is the cross-product of the associated vector fields
Lemma 4.2.8 The wedge product of a one-form and a two-form is the dot product of the associated vect
Lemma 4.3.2 The exterior derivative in R3
Lemma 4.3.6 The graded product rule for the exterior derivative
Lemma 4.3.9 d2 = 0
Lemma 4.4.9 Vector calculus identities, part 1
Lemma 4.4.10 Vector calculus identities, part 2
Lemma 4.4.11 Vector calculus identities, part 3
Lemma 4.4.12 Vector calculus identities, part 4
Lemma 4.6.3 Exact k-forms are closed
Theorem 4.6.4 Poincare’s lemma for k-forms, version 1
Theorem 4.6.5 Poincare’s lemma for k-forms, version II
Lemma 4.7.1 The pullback of a k-form
Lemma 4.7.4 The pullback commutes with the exterior derivative
Lemma 4.7.7 The pullback of a top form in Rn in terms of the Jacobian determinant
Theorem 4.7.8 Inverse Function Theorem
Theorem 4.7.9 Implicit Function Theorem
Lemma 4.7.11 An explicit formula for the pullback of a basic one-form
Lemma 4.7.13 The pullback commutes with the wedge product
Corollary 4.7.14 An explicit formula for the pullback of a basic k-form
Lemma 4.7.15 The pullback of a basic n-form in Rn
Lemma 4.8.8 The Laplace-Beltrami operator and the Laplacian of a function
Lemma 4.8.9 The Laplace-Beltrami operator and the Laplacian of a vector field
Lemma 4.8.10 Vector calculus identities, part 5
List of definitions
297
APPENDIX B. LIST OF DEFINITIONS 298
Chapter 4 k-forms
Definition 4.1.1 The basic one-forms
Definition 4.1.3 Basic two-forms
Definition 4.1.4 Basic three-forms
Definition 4.1.5 Basic k-forms
Definition 4.1.8 k-forms in R3
Definition 4.2.1 The wedge product
Definition 4.3.1 The exterior derivative of a k-form
Definition 4.4.1 The gradient of a function
Definition 4.4.2 The curl of a vector field
Definition 4.4.3 The divergence of a vector field
Definition 4.6.1 Exact and closed k-forms
Definition 4.7.5 The Jacobian
Definition 4.7.6 Top form
Definition 4.7.10 The pullback of a basic one-form with respect to a linear map
Definition 4.7.12 The pullback of a basic k-form with respect to a linear map
Definition 4.8.1 The Hodge star dual of a k-form in Rn
Definition 4.8.7 The codifferential and the Laplace-Beltrami operator
List of examples
300
APPENDIX C. LIST OF EXAMPLES 301
Chapter 4 k-forms
Example 4.2.2 The wedge product of two one-forms
Example 4.2.3 The wedge product of a one-form and a two-form
Example 4.2.4 The wedge product of a zero-form and a k-form
Example 4.3.3 The exterior derivative of a zero-form on R3
Example 4.3.4 The exterior derivative of a one-form on R3
Example 4.3.5 The exterior derivative of a two-form on R3
Example 4.3.7 The exterior derivative of the wedge product of two one-forms
Example 4.4.7 Maxwell’s equations
Example 4.5.1 The direction of steepest slope
Example 4.5.4 The curl of the velocity field of a moving fluid
Example 4.5.5 An irrotational velocity field
Example 4.5.6 Another irrotational velocity field
Example 4.5.7 The divergence of the velocity field of an expanding fluid
Example 4.5.8 An imcompressible velocity field
Example 4.5.9 Another incompressible velocity field
Example 4.6.2 Exact and closed one-forms in R3
Example 4.7.2 The pullback of a two-form
Example 4.7.3 The pullback of a three-form
Example 4.8.2 The action of the Hodge star in R
Example 4.8.3 The action of the Hodge star in R2
Example 4.8.4 The action of the Hodge star in R3
Example 4.8.5 An example of the Hodge star action in R3
Example 4.8.6 Maxwell’s equations using differential forms (optional)
Example 5.3.3 Integral of a two-form over a rectangular region with canonical orientation
Example 5.3.4 Integral of a two-form over an x-supported (or type I) region with canonical orientation
Example 5.3.8 Area of a disk
Example 5.4.4 The graph of a function in R3
Example 5.4.5 The sphere
Example 5.4.6 The cylinder
Example 5.4.7 Grid curves on the sphere
Example 5.5.6 Upper half-sphere
Example 5.6.2 An example of a surface integral
Example 5.6.7 An example of a surface integral of a vector field
Example 5.7.4 Using Green’s theorem to calculate line integrals
Example 5.7.5 Area of an ellipse
Example 5.8.5 Using Stokes’ theorem to evaluate a surface integral by transforming it into a line integral
Example 5.8.7 Using Stokes’ theorem to evaluate a surface integral by using a simpler surface
Example 5.8.8 Using Stokes’ theorem to evaluate a line integral by transforming it into a surface integral
Example 5.9.2 The electric flux and net charge of a point source
List of exercises
303
APPENDIX D. LIST OF EXERCISES 304
Chapter 4 k-forms
Exercise 4.1.5.1
Exercise 4.1.5.2
Exercise 4.1.5.3
Exercise 4.1.5.4
Exercise 4.1.5.5
Exercise 4.2.3.1
Exercise 4.2.3.2
Exercise 4.2.3.3
Exercise 4.2.3.4
Exercise 4.2.3.5
(Continued on next page)
APPENDIX D. LIST OF EXERCISES 305
Exercise 4.2.3.6
Exercise 4.3.4.1
Exercise 4.3.4.2
Exercise 4.3.4.3
Exercise 4.3.4.4
Exercise 4.3.4.5
Exercise 4.3.4.6
Exercise 4.3.4.7
Exercise 4.3.4.8
Exercise 4.3.4.9
Exercise 4.4.5.1
Exercise 4.4.5.2
Exercise 4.4.5.3
Exercise 4.4.5.4
Exercise 4.4.5.5
Exercise 4.4.5.6
Exercise 4.4.5.7
Exercise 4.5.4.1
Exercise 4.5.4.2
Exercise 4.5.4.3
Exercise 4.5.4.4
Exercise 4.5.4.5
Exercise 4.5.4.6
Exercise 4.6.3.1
Exercise 4.6.3.2
Exercise 4.6.3.3
Exercise 4.6.3.4
Exercise 4.6.3.5
Exercise 4.6.3.6
Exercise 4.7.5.1
Exercise 4.7.5.2
Exercise 4.7.5.3
Exercise 4.7.5.4
Exercise 4.7.5.5
Exercise 4.7.5.6
Exercise 4.7.5.7
Exercise 4.8.4.1
Exercise 4.8.4.2
Exercise 4.8.4.3
Exercise 4.8.4.4
Exercise 4.8.4.5
(Continued on next page)
APPENDIX D. LIST OF EXERCISES 306
Exercise 4.8.4.6
Exercise 5.8.3.4
Exercise 5.8.3.5
Exercise 5.8.3.6
Exercise 5.9.3.1
Exercise 5.9.3.2
Exercise 5.9.3.3
Exercise 5.9.3.4