Manifolds 2014 Parallelization
Manifolds 2014 Parallelization
Ed Segal
Autumn 2014
Contents
1 Introduction
manifolds
4
. . . . . . . . . . . . . . . . 4
. . . . . . . . . . . . . . . . 7
. . . . . . . . . . . . . . . . 12
3 Submanifolds
17
3.1 Definition of a submanifold . . . . . . . . . . . . . . . . . . . . 17
3.2 Level sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4 Smooth functions
27
4.1 Definition of a smooth function . . . . . . . . . . . . . . . . . 27
4.2 The rank of a smooth function . . . . . . . . . . . . . . . . . . 33
4.3 Some special kinds of smooth functions . . . . . . . . . . . . . 37
5 Tangent spaces
5.1 Tangent vectors via curves . . . . . .
5.2 Tangent spaces to submanifolds . . .
5.3 A second definition of tangent vectors
5.4 A third definition of tangent vectors .
6 Vector fields
6.1 The tangent bundle . . . . . . .
6.2 Vector fields and flows . . . . .
6.3 Other definitions of vector fields
6.4 Vector bundles . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
40
41
47
50
52
.
.
.
.
60
60
63
69
77
125
127
Introduction
Figure 1: A torus.
Lets go down a dimension:
Example 1.3. A circle, sometimes denoted S 1 , is an example of a 1-dimensional
manifold. A small piece of a circle looks just like a small piece of the real
line R.
Heres an example of a different flavour:
Example 1.4. Let Mat22 (R) be the set of all 2 2 real matrices, this is a
4-dimensional real vector space so its isomorphic to R4 . Now let
GL2 (R) Mat22 (R)
denote the subset of invertible matrices. If M is an invertible matrix, and N
is any matrix whose entries are sufficiently small numbers, then the matrix
M +N will still be invertible. So every matrix nearby M also lies in GL2 (R)
(i.e. GL2 (R) is an open subset). This means that a small neighbourhood of
M looks exactly like a small neighbourhood of the origin in Mat22 (R)
= R4 .
Hence GL2 (R) is an example of a 4-dimensional manifold.
This is an example of a Lie group, a group that is also a manifold. Lie
groups are very important, but they wont really be covered in this course.
We often picture manifolds as being subsets of some larger vector space,
e.g. we think of S 2 or T 2 as smooth surfaces sitting inside R3 . This is very
helpful for our intuition, but the theory becomes much more powerful when
we can talk about manifolds abstractly, without reference to any ambient
vector space. A lot of the hard work in this course will involve developing
the necessary machinery so that we can do this.
3
2.1
Topological manifolds
f : U U
f : U U Rn
with x U .
In words, this says that at any point in X we can find an open neighbourhood which is homeomorphic to some open set in Rn . A concise (but
slightly imprecise) way to say this is X is locally homeomorphic to Rn .
Example 2.3. The circle S 1 is a 1-dimensional topological manifold. Lets
prove this carefully. Firstly, lets define S 1 to be the subset
S 1 = (x, y); x2 + y 2 = 1 R2
and equip it with the subspace topology. Next we need to find some coordinate charts, well do this using stereographic projection.
4
Let (x, y) be a point in S 1 , not equal to (0, 1). Draw a straight line
through (x, y) and the point (0, 1), and let x R be the point where this
line crosses the x-axis, so:
x
x =
1+y
f2 : U2 U2
x
(x, y) 7
1y
U2 = Rn
and:
f2 : U2 U2
x0
xn1
(x0 , ..., xn ) 7
, ...,
1 xn
1 xn
Since every point in S n lies in at least one of U1 and U2 , this proves that S n
is a topological manifold, of dimension n.
2.2
Smooth atlases
f1 : U1 U2 R \ 0
Similarly, the point (0, 1) lies in the co-ordinate chart U2 , and it gets
mapped by f2 to the origin in R. So the function f2 also defines a homeomorphism:
f2 : U1 U2 R \ 0
To change between co-ordinates we must consider the composition:
21 = f2 f11 : R \ 0 R \ 0
This sends x R \ 0 to:
21 (
x) = f 2
1 x2
2
x
,
1 + x2 1 + x2
=
1
x
So if our point has co-ordinate x under our first co-ordinate chart, then it has
co-ordinate 1/
x in our second chart. The function 21 is called a transition
function.
7
f1 : U1 U1 Rn
and
f2 : U2 U2 Rn
f1 : U1 U2 f1 (U1 U2 ) U1
The image of this homeomorphism is some open subset of Rn , contained in
U1 . Similarly, f2 gives us a homeomorphism
f2 : U1 U2 f2 (U1 U2 ) U2
f2 (U1 U2 ) = Rn \ 0
(in this example both f1 (U1 U2 ) and f2 (U1 U2 ) happen to be the same
subset of Rn , but this is a coincidence). The transition function between
these two charts is the function:
21 : Rn \ 0 Rn \ 0
x0
xn1
(
x0 , ..., xn1 ) 7 P 2 , ..., P 2
xi
xi
A transition function tells us how to change co-ordinates between two
different charts in some region of our manifold, so it tells us each new coordinate as some continous function of the old co-ordinates. However, we
dont want our change-of-co-ordinate functions to be merely continuous, we
really want them to be smooth.
Recall that a function
F : Rn Rm
is called smooth (or C ) if we can take partial derivatives of F to any order,
in any direction. This definition also makes sense if F is only defined on some
open subset of Rn . Since a transition function is a function from an open set
in Rn to some other open set in Rn , it makes sense to ask if the transition
function is smooth.
Definition 2.7. Let X be a topological manifold. An atlas for X is a
collection of co-ordinate charts on X
fi : Ui Ui Rn ,
iI
ij : fj (Ui Uj ) fi (Ui Uj ),
is a smooth function.
9
i, j I
f0 : U0 U0
be the identity function. This is a co-ordinate chart, and since it covers all of
X it is in fact an atlas (consisting of a single chart). Hence Rn is a topological
manifold, of dimension n. Furthermore this is a smooth atlas, because there
are no non-trivial transition functions!
More generally we can let X be any open set inside Rn , then the same
procedure provides a smooth atlas on X (with a single chart).
2.3
Smooth structures
Weve just seen (in Example 2.12) that there is a trivial smooth atlas on the
topological manifold X = R. Here is another smooth atlas on this topological
manifold:
Example 2.13. Let X = R. Here are two co-ordinate charts on X:
1
U1 = R>0 , U1 = R>0 , f1 (x) =
x
1
U2 = R<1 , U2 = R>0 , f2 (x) =
1x
The union of U1 and U2 is the whole of R, so together they form an atlas.
Lets look at the transition functions; we have
U1 U2 = (0, 1),
f1 (U1 U2 ) = R>1 ,
f2 (U1 U2 ) = R>1
12
10 = f1 : R>0 R>0
and
20 = f2 : R<1 R>0
and their inverses. Since all four of f1 , f2 , f11 and f21 are smooth, this shows
that both charts in B are compatible with the atlas A (and the chart in A
is compatible with the atlas B). Hence these two atlases are compatible.
13
As weve said, knowing exactly which charts are in our atlas A is not so
important, all we really care about is the set of charts that are compatible
with A. The next lemma says that if we replace A by a compatible atlas B,
then this information does not change: the set of compatible charts remains
the same.
Lemma 2.17. Let X be a topological manifold, and let
A = {(Ui , fi ); i I}
B = {(Uj , fj ); j J}
and
j0 : f0 (U0 Uj ) fj (U0 Uj )
We need to show that both j0 and 1
j0 are smooth functions. Pick any point
U0 Uj . Were going to show that j0 is smooth at the point f0 (x), and that
1
j0 is smooth at the point fj (x) = j0 (x). If we can do this for any point
x, then well have shown that both functions are smooth, and proved the
lemma.
Since A is an atlas, there exists some chart (Ui , fi ) A with x Ui . Set
W = U0 Uj Ui , this is an open neighbourhood of x. We have homeomorphisms:
W
f0
f0 (W )
fj
fi
fi (W )
fj (W )
fj f01 : f0 (W ) fj (W )
is just the restriction of the transition function j0 . Similar statements apply
when we move between fi (W ) and either of the other two charts, so we have
that:
j0 |f0 (W ) = ji |fi (W ) i0 |f0 (W )
Since ji and i0 are smooth by assumption, it follows that j0 is smooth
within the open set f0 (W ), and in particular it is smooth at the point f0 (x).
1
1
Since 1
ij and i0 are also smooth, j0 is smooth at the point fj (x).
14
16
Submanifolds
3.1
Definition of a submanifold
f : U U Rn
U = R>0 (, )
and:
f 1 : U U
(r, ) 7 (r cos , r sin )
Then its clear that f 1 is a smooth bijection. Its slightly less obvious that
the inverse function f : U U is also smooth, but this can be shown using
the Inverse Function Theorem (more on this shortly). Hence (U, f ) is a
co-ordinate chart on R2 , compatible with the standard atlas. We have:
f (S 1 U ) = {(1, )} U
Now consider the map:
: R2 R2
(r, ) 7 (, r 1)
This is an affine map (i.e. the composition of a linear map and a translation),
its invertible, smooth, and has a smooth inverse. So if we compose f with
then we get a new co-ordinate chart
f : U U = (, ) R>1
18
and we have:
( f )(S 1 U ) = {(, 0)} = R U
Every point on S 1 , apart from the point (1, 0), lies in U1 , so this co-ordinate
chart demonstrates that S 1 satisfies the submanifold condition at every point
except for (1, 0). To deal with this final point we need a second co-ordinate
chart, we can do this by using polar co-ordinates with (0, 2) (so we
delete the positive x-axis).
As the previous example shows, its enough to find a co-ordinate chart
such that f (U Z) is the intersection of U with some m-dimensional affine
subspace of Rn . This is because we can always apply an affine change-of-coordinates to turn it into our standard subspace Rm Rn .
As the name suggests, an m-dimensional submanifold is itself a manifold,
of dimension m.
Proposition 3.4. Let X be an n-dimensional manifold, and let Z X be
an m-dimensional submanifold of X. Then we can equip Z with the structure
of an m-dimensional manifold.
Proof. Firstly we equip Z with the subspace topology to make it a topological
space. Now consider a co-ordinate chart
f : U U Rn
on X having the property that:
f (Z U ) = Rm U
(3.5)
The intersection V = U Z is an open set in Z, and f induces a homeomorphism between V and the open set:
V = Rm U
Hence (V, f ) is a co-ordinate chart on Z. Let A0 be the set of all co-ordinate
charts on X having the property (3.5) (this is a hugely infinite set!), this
gives a corresponding set A of co-ordinate charts on Z. By the definition of
a submanifold, every point in Z lies in at least one of the charts in A, so A
is an atlas.
Now lets look at the transition functions. Pick two charts (U1 , f1 ),
(U2 , f2 ) in A0 , and let (V1 , f1 ), (V2 , f2 ) be the corresponding charts in A.
Note that V1 V2 = Z (U1 U2 ), so
f1 (V1 V2 ) = Rm f1 (U1 U2 )
19
and:
f2 (V1 V2 ) = Rm f2 (U1 U2 )
We have a smooth transition function
21 : f1 (U1 U2 ) f2 (U1 U2 )
between the charts on X, and restricting this to the subspace Rm gives the
transition function
21 : f1 (V1 V2 ) f2 (V1 V2 )
3.2
Level sets
and h3 (x, y) = x2 + y 2
20
and:
f : U U
(x, y) 7 (x, y /x)
Note that this really is a co-ordinate chart (f is smooth with a smooth
inverse), and that:
f (Z U ) = {(x, 0)} U
Since Z is entirely contained in U (provided that 6= 0), this demonstrates
that Z is a submanifold.
However, for = 0, the level set
Z0 = {xy = 0} = {x = 0} {y = 0}
consists of both co-ordinate axes. This certainly doesnt look like a submanifold, because of the singularity at the point where the two axes cross. In
fact one can prove that Z0 is not even a topological manifold.
So we need to find an additional condition to guarantee that a level set
of h is a submanifold. To do this we need to look at the derivative of h.
For a smooth function
h = (h1 , ...., hk ) : Rn Rk
and a point x = (x1 , ..., xn ) Rn , recall that the derivative of h at x is the
linear map
Dh|x : Rn Rk
given by the k-by-n matrix
Dh|x =
hi
xj x
22
value). Hopefully this makes intuitive sense: we start with a space having n
degrees-of-freedom, then we impose k equations, so we have n k degreesof-freedom left.
The proposition will follow fairly easily from the following important theorem, which you should recall from real analysis.
Theorem 3.12 (Inverse Function Theorem). Let W1 and W2 be two open
subsets of Rn , and let
F : W1 W2
be a smooth function. Let x W1 be a point such that the derivative of F at
x
DF |x : Rn Rn
is an isomorphism. Then there exists an open neighbourhood of U W1 of
x such that the function
F : U F (U )
W2
hk
xj
x
This collection of n vectors v1 , ..., vn together span Rk (since Dh|x is a surjection), so by elementary linear algebra there must be some subset of them
which forms a basis for Rk . After re-ordering the co-ordinates xj if necessary,
we may assume that the subset vnk+1 , ..., vn is a basis, i.e. the k-by-k matrix
h1
h1
...
xn
xnk+1 x
x
..
(3.13)
M =
: Rk Rk
.
hk
hk
... x
xnk+1
n
x
is an isomorphism.
Now consider the function:
f : Rn 7 Rn
x 7 (x1 , ..., xnk , h1 (x), ..., hk (x))
The derivative of f at our point x Z0 is an n-by-n matrix of the form
Ink 0
Df |x =
?
M
where M is the matrix (3.13), and Ink is the identity matrix. Hence
det Df |x = det M , and this is non-zero since M is an isomorphism. Applying the Inverse Function Theorem (Theorem 3.12), we see that there is
an open set U Rn containing x such that the function
f : U f (U ) Rn
is a bijection with a smooth inverse. This is our required co-ordinate chart.
25
Intuitively, critical points and critical values are rather rare. If we pick
a point x Rn at random, then it is vanishingly unlikely that the derivative Dh|x is not surjective, since almost all k-by-n matrices are surjective
(provided that k n). This suggests that almost all level sets of a smooth
function are submanifolds. This intuition is correct, and can be turned into
a result known as Sards Theorem. However it would take us on a significant
detour to even state this result precisely!
Proposition 3.11 gives us an easy way to find new manifolds: just pick
any smooth function from Rn to Rk and then look at a level set. Provided
that were at a regular value (and by Sards Theorem we almost always will
be), the level set is a manifold of dimension n k.
Example 3.14. Consider the smooth function
h : Rn R
h : (x1 , ..., xn ) 7 x21 + ... + x2n
The derivative of h at a point (x1 , ..., xn ) is the 1-by-n matrix
(2x1 , ...., 2xn ) : Rn R
which is a surjection if at least one of the xi is not zero. Hence the origin is
the only critical point of h, and 0 R is the only critical value. For R>0 ,
the level set Z = h1 () is a (n 1)-dimensional sphere of radius . So
this sphere is a codimension-1 submanifold of Rn , and in particular it is a
(n 1)-dimensional manifold.
If R<0 then Z = (the empty set), technically this is also a
codimension-1 submanifold of Rn , but its less interesting!
Weve now seen two ways to get an atlas on the n-dimensional sphere S n :
in Example 2.4 we used stereographic projection to get an atlas with two
charts, or since S n is a submanifold of Rn+1 we can use Proposition 3.4 to
get an atlas with infinitely-many charts. In fact these two atlases define the
same smooth structure on S n (i.e. theyre compatible), but we wont fill in
all the details of this fact.
We can be slightly more general. Take an open set
X Rn
and consider a smooth function
h : X Rk
26
4
4.1
Smooth functions
Definition of a smooth function
f : U U Rn
on X then we can consider the function
= h f 1 : U R
h
as the function h written in this choice of co-ordinates.
We should think of h
is a smooth function, then we should declare that h is also smooth, at
If h
least within the open set U X. If we want to be more specific then we can
choose a particular point x U , and declare that h is smooth at x iff the
is smooth at the point f (x).
function h
However, there might be a problem with this definition: it might depend
on which co-ordinates we chose! Lets check that it doesnt. Suppose that
(U1 , f1 ) and (U2 , f2 ) are two co-ordinate charts on X, both containing the
point x. Then the functions
1 = h f 1
h
1
and
2 = h f 1
h
2
f11
: x 7
2
x
1 + x2
2
and this is a smooth function. We can similarly check that the function
h f21 is smooth, where (U2 , f2 ) is the other co-ordinate chart.
Obviously we could replace h with any other smooth function of x and y
in this example, and it would again define a smooth function on S 1 .
If X = Rn (or an open set in Rn ) equipped with the standard smooth
structure, its clear that this definition of a smooth function agrees with the
ordinary definition.
Its obvious how to generalize Definition 4.1 to get a definition of when a
function
h : X Rk
29
f : U U Rn
with x U , and pick a co-ordinate chart on Y
g : V V Rk
f : UH,V f (UH,V ) Rn
Then H defines a function
H : UH,V V
and we can consider the composition:
g H f 1 : f (UH,V ) V
This is a function between an open set in Rn and an open set in Rk , so it
makes sense to ask if it is a smooth function.
30
f : U U Rn
g : V V Rk
on Y containing the point H(x), the function
g H f 1 : f (UH,V ) V
is smooth.
We say that H is smooth everywhere, or just smooth, iff H is smooth
at x for every point x X.
Notice that a continuous function H is smooth iff, for any co-ordinate
chart (U, f ) on X, and any co-ordinate chart (V, g) on Y , the composition
g H f 1 : f (UH,V ) V
is a smooth function.
If we set Y = R (with the standard smooth structure), then its easy to
check that this definition agrees with Definition 4.1. Recall that in that case
we observed that if a function looks smooth in one co-ordinate chart then
it looks smooth in all co-ordinate charts. This is still true for our more
general definition of a smooth function. Suppose (U1 , f1 ) and (U2 , f2 ) are
two co-ordinate charts on X, and let 21 be the transition function between
them. Now let (V1 , g1 ) and (V2 , g2 ) be two co-ordinate charts on Y , and let
21 be the transition function between them. Then we have an equality of
functions
g2 H f21 = 21 (g1 H f11 ) 12
(4.4)
(on the open subset in U2 where both sides are defined). Since all the transition functions are smooth, the function g2 H f21 will be smooth iff the
function g1 H f11 is smooth.
In particular if we want to check that H is a smooth function, we can
pick an atlas {(Ui , fi )} for X and an atlas {(Vj , gj )} for Y , and then check
that every function gj H fi1 is smooth.
31
Example 4.5. Let X = T 1 (from Example 2.10) and let Y = S 1 . Now let:
H : T 1 S1
[t] 7 (cos 2t, sin 2t)
(note that this is well-defined). Lets show that H is smooth.
Take the co-ordinate chart (U1 , f1 ) on T 1 , where U1 = T 1 \ [0] and
f1 : U1 U1 = (0, 1) R
x
1+y
and
G:Y Z
the function h G g 1 is a smooth function, defined on some open neighbourhood of the point g(H(x)) Rk . To prove that G H is smooth at x,
we need to know that the function
h G H f 1 : f (UGH,W ) W
is smooth at the point f (x). But in a sufficiently small open neighbourhood
of f (x) we can factor this function as
h G H f 1 = (h G g 1 ) (g H f 1 )
and both factors are smooth.
If you know what a category is, then this shows that there is a category
whose objects are manifolds and whose arrows are smooth functions.
4.2
and
G : Rk Rm
(4.8)
So the linear maps DF1 |f1 (x) and DF2 |f2 (x) are not the same, but they are related by this formula. Now we can make an important observation: the linear
34
maps D21 |g1 (H(x)) and D12 |f2 (x) are isomorphisms (because the transition
functions are bijections with smooth inverses), this means that the rank of
DF2 |f2 (x) must be the same as the rank of DF1 |f1 (x) . Consequently we can
make the following definition:
Definition 4.9. Let X and Y be manifolds (of dimensions n and k respectively) and let F : X Y be a smooth function. Fix a point x X. Now
pick a co-ordinate chart (U, f ) containing x and a co-ordinate chart (V, g)
containing F (x), and consider the function:
F = g F f 1 : UF,V V
We define the rank of F at x to be the rank of the derivative
DF |f (x) : Rn Rk
of F at f (x).
This makes sense because of the formula (4.8); it doesnt matter which
co-ordinate charts we choose, the rank of DF |f (x) will always be the same.
Now we can generalize Definitions 3.8 and 3.9.
Definition 4.10. Let F : X Y be a smooth function between two manifolds, of dimensions n and k respectively. We say that a point x X is a
regular point of F if the rank of F at x is equal to k. If x is not a regular
point then we call it a critical point.
We say that a point y Y is a regular value of F if every point x
F 1 (y) is a regular point. If y is not a regular value then we call it a critical
value.
So x is a regular point of F iff the derivative
DF |f (x) : Rn Rk
is a surjection, where F is F written in any co-ordinate charts. In other
words x is a regular point of F iff f (x) is a regular point of F , for any choice
of co-ordinates. Clearly if we set X to be an open set in Rn , and Y to be
Rk , then we recover our previous definitions.
We can also generalize Proposition 3.11 fairly easily:
Proposition 4.11. Let F : X Y be a smooth function between two manifolds, of dimensions n and k respectively. Let y Y be a regular value of F .
Then the level set
Zy = F 1 (y) X
is a submanifold of X of codimension k.
35
Rn
h : W W
f11 : R2 U1
2
x
2
y
1 x2 y2
(
x, y) 7
,
,
1 + x2 + y2 1 + x2 + y2 1 + x2 + y2
In this chart, the function F becomes the function
F = F f11 : (
x, y) 7
2
x
2
1 + x + y2
DF |(x,y) =
,
: R2 R
(1 + x2 + y2 )2 (1 + x2 + y2 )2
This has rank 1 except when it goes to zero, which occurs exactly at (
x, y) =
(1, 0). Hence inside the open set U1 , the only critical points of F are:
f1 (1, 0) = (1, 0, 0) S 2
36
and we can use another chart to check that these are really the only critical
points in S 2 . Thus the critical values of F are = 1 and = 1.
If || < 1 then the level set F 1 () is a circle, the intersection of the 2sphere with the plane {x = }. The theorem says that this is a 1-dimensional
submanifold of S 2 .
If || > 1 then the level set F 1 () is empty. At the critical values = 1
the level set consists of a single point, this is evidently not a 1-dimensional
submanifold.
4.3
We can use our definition of rank to single out some particularly important
kinds of smooth functions.
Definition 4.13. A smooth function F : X Y is called a submersion if
the rank of F at any point is equal to the dimension of Y .
So F is a submersion iff the derivative at any point (in any co-ordinates)
is a surjection, i.e. a submersion is exactly a smooth function that has no
critical points. There is a dual notion to this:
Definition 4.14. A smooth function F : X Y is called an immersion if
the rank of F at any point x X is equal to the dimension of X.
In other words, F is an immersion iff the derivative at any point (in any
co-ordinates) is an injection.
Now let X be a manifold, and let Z X be a submanifold of X. Recall
from Proposition 3.4 that there is an induced smooth structure on Z, making
it into a manifold in its own right.
Lemma 4.15. Let X be a manifold, and let Z X be a submanifold of X.
Then the inclusion map
: Z , X
is smooth, and an immersion.
Proof. Exercise.
However, not every immersion is of this form.
Example 4.16. Let X = R and Y = R2 , and let:
F : R R2
t 7 (t2 , t3 t)
37
If two manifolds are diffeomorphic then they are exactly the same, for all
practical purposes (it may help to think of diffeomorphic as another word
for isomorphic).
Suppose F : X Y is a diffeomorphism, and we pick a point x X, a coordinate chart (U, f ) containing x, and a co-ordinate chart (V, g) containing
F (x). By shrinking U and V is necessary, we can assume that F is a bijection
from U to V and hence
F = f F g 1 : U V
is a smooth bijection with a smooth inverse. This means that the derivative
of F at any point must be an isomorphism, so F is both a submersion and
an immersion. In particular, diffeomorphic manifolds must have the same
dimension, which is reassuring!
There do exist smooth functions which are bijections, but whose inverse
functions are not smooth. These functions are not diffeomorphisms. This
means the following criterion is useful:
Lemma 4.18. Let X and Y be n-dimensional manifolds, and let
F :XY
be a smooth bijection. If the rank of F is n at every point then F is a
diffeomorphism.
Proof. We just need to show that the inverse function F 1 is smooth. Fix a
point y Y , let x = F 1 (y), and choose co-ordinate charts (U, f ) containing
x and (V, g) containing y. Assume (by shrinking U if necessary) that F (U )
V , and consider the function F = g F f 1 . Since the rank of F is n at
the point F (y), the derivative
DF |f (x) : Rn Rn
is an isomorphism. By the Inverse Function Theorem, there is some open
neighbourhood of g(y) on which the function F 1 is smooth. This proves
that F 1 is smooth at y.
Example 4.19. Recall from Example 4.5 that we have a smooth function:
H : T 1 S1
[t] 7 (cos 2t, sin 2t)
39
2
:RR
1 + sin 2t
This is never zero, which shows that H has rank 1 at all points other than
[0], [ 34 ] T 1 . We can check using other charts that H also has rank 1 at these
remaining points, so by Lemma 4.18 H is a diffeomorphism.
So our two versions of the circle, T 1 and S 1 , are diffeomorphic manifolds.
Example 4.20. Recall from Example 2.23 that we can find a non-standard
atlas C on the topological manifold R which is not compatible with the standard atlas A. However, the two smooth manifolds (R, [A]) and (R, [C]) are
diffeomorphic (we leave the proof as an exercise).
This leaves an interesting question: does there exist any smooth atlas D
on Rn such that the resulting smooth manifold (Rn , [D]) is not diffeomorphic to the standard Rn ? This question was comprehensively answered in
the 1980s, and the answer is one of the most astonishing results in all of
mathematics!
Tangent spaces
In fact defining tangent spaces is the hard bit, the fact that we can define
DF |x will follow almost automatically.
5.1
(5.1)
=f
is a curve through the point x = f (x) U (we might have to shrink to
ensure that the image of lies within U ). Hence this curve has an associated
tangent vector:
D
|0 Rn
If we have two curves through x, say and , then we could say that they
are tangent at x if when we look at them in co-ordinates then the two curves
are the same. This is a reasonable definition, but we need to check that it
doesnt depend on which co-ordinates we chose.
So, pick two different charts (U1 , f1 ) and (U2 , f2 ) both containing x. Looking at the curve in these two charts gives a curve
1 = f1 through the
point x1 = f1 (x), and a curve
2 = f2 through the point x2 = f2 (x). These
curves are related by the transition function 21 between the two charts:
2 = 21
1
(we can always shrink to make sure that lands in U1 U2 ). If we use our
first chart, the tangent vector to would be the vector D
1 |0 Rn , and if
we use our second chart, it would be D
2 |0 . By the chain rule, we have that
D
2 |0 = D21 |x1 D
1 |0
(5.3)
so these two vectors are related by the linear map:
D21 |x! : Rn Rn
This means that our definition of when two curves through x are tangent is
indeed independent of co-ordinates. Suppose we have two curves and
through x, and they have the same tangent vector when we look at them in
the chart (U1 , f1 ), i.e.
D
1 |0 = D
1 |0
Then by equation (5.3), we must also have
D
2 |0 = D
2 |0
so and have the same tangent vector when we look at them in the chart
(U2 , f2 ). Lets write down our definition formally:
Definition 5.4. Fix a point x in a manifold X. We say that two curves ,
through x are tangent at x iff for any co-ordinate chart (U, f ) containing
x, we have
D(f )|0 = D(f )|0
As we have shown, if this holds in one chart then it holds in all charts.
Obviously being tangent at x is an equivalence relation on the set of all
curves through x.
Definition 5.5 (Geometers definition ). Fix a point x in a manifold X. A
tangent vector to x is an equivalence class of curves through x. We denote
the set of all tangent vectors to x by
Tx X = {curves through x} /(tangency at x)
and call it the tangent space to X at x.
43
(5.7)
v : (, ) U
t 7 f (x) + vt
for small-enough (as in (5.1)), and then v = f 1
v is a curve through x
such that f (v ) = v. Hence f is a bijection of sets.
We can use the bijection f to put a vector space structure on Tx X,
i.e. we can define an addition operation
[] + [ ] = 1
f () + f ( )
(5.8)
f
and a scalar multiplication
[] = 1
()
f
f
(5.9)
(and these are guaranteed to satisfy the vector space axioms). However, once
again we need to check that these definitions dont depend on our choice of
co-ordinates.
So, pick two co-ordinate charts (U1 , f1 ) and (U2 , f2 ) both containing x.
This gives two different bijections
f1 : Tx X Rn
[] 7 D(f1 )|0
and
f2 : Tx X Rn
[] 7 D(f2 )|0
given by calculating tangent vectors in the two different charts. By the chain
rule (5.3), we have that
D(f2 )|0 = D21 |f1 (x) D(f1 )|0
44
where 21 is the transition function between our two charts. So the two
bijections f1 and f2 are related by:
f2 = D21 |f1 (x) f1
(5.10)
Since D21 is a linear isomorphism, this implies that the operations (5.8)
and (5.9) give the same result, independent of which chart we used to define
them.
So we have achieved our first aim for this section, namely to any point
x in a manifold X we have attached a vector space Tx X. If we choose any
co-ordinate chart (U, f ) containing x, then we get a linear isomorphism
f : Tx X Rn
as in (5.7). However, there is no canonical way to identify Tx X with Rn , and
indeed if we have two different charts then we know that f1 and f2 are
related by the equation (5.10).
Now we move on to our second aim: defining the derivative of a smooth
function between two manifolds.
Firstly, suppose that U and V are open sets in Rn and Rm respectively,
and that F is a smooth function:
F : U V
Pick a point x U and let y = F (
x). If we have a curve through x, then
the compostion F is a curve through y (since the composition of two
smooth functions is smooth). Furthermore, using the chain rule again tells
us that:
D(F )|0 = DF |x D|0
(5.11)
In particular, the tangent vector to F only depends on the tangent vector
to . So if we wish, we could view the derivative DF |x as a function:
DF |x :
{curves through y}
{curves through x}
(tangency at x)
(tangency at y)
[] 7 [F ]
defined by:
DF |x () = [F ] Ty Y
We call this linear map the derivative of F at x.
Proof. Pick a chart (U, f ) on X containing x, and a chart (V, g) on Y containing y. In these charts, F becomes the function:
F = g F f 1 : U V
Now choose a curve through x. Using our chart, this becomes a curve
through the point y = g(y) V . The tangent vector associated to this curve
is
g (F ) = D(F
)|0 Rm
which by the chain rule (5.11) is equal to:
DF |x D
|0 = DF |x f ()
So the tangent vector g (F ) only depends on the tangent vector f ().
This means that the equivalence class of the curve F in the space Ty Y only
depends on the equivalence class of the curve in the space Tx X, and thus
we have a well-defined function DF |x : Tx X Ty Y that sends [] [F ].
Furthermore the square
Tx X
DF |x
Rn
Ty Y
g
DF |x
(5.13)
Rm
commutes, i.e.
f
DF |x = 1
g D F |x
and therefore DF |x is a linear map, since its the composition of three linear
maps.
46
5.2
)|(1,) =
cos sin
sin cos
so the tangent space to the point (cos , sin ) S 1 is the line spanned by
the vector ( sin , cos )> R2 .
49
5.3
Were now going to discuss a second way to define tangent vectors, and later
on well introduce a third definition. These other definitions are precisely
equivalent to the definition weve already introduced, but each one has its
own advantages and disadvantages.
Fix a point x in a manifold X, and fix a tangent vector [] Tx X, the
equivalence class of some curve through x. If we now choose a co-ordinate
chart (U, f ) containing x, we can turn [] into an ordinary column vector
f () Rn . Furthermore, if we change co-ordinates between (U1 , f1 ) and
(U2 , f2 ), we know that the transformation law
f2 () = D21 |f1 (x) f1 ()
(5.18)
holds (this was equation (5.10)). If we wish, we can take these properties to
be the definition of a tangent vector.
Definition 5.19 (Physicists definition ). Fix a point x in a manifold X.
Let Ax denote the set of all co-ordinate charts on X that contain the point
x. A tangent vector to x is a function
: Ax Rn
which we write
: (U, f ) 7 f
and which has the following property: for any two charts (U1 , f1 ) and (U2 , f2 )
in Ax , the equation
f2 = D21 |f1 (x) f1
(5.20)
holds.
Lets (temporarily) use the notation Tx X to denote the set of tangent
vectors in the sense of Definition 5.19, although as we shall see in a moment
the two definitions are equivalent and Tx X is the same thing as Tx X.
Lemma 5.21. For any chart (U, f ) Ax , the function evaluate in (U, f ):
evf : Tx X Rn
7 f
is a bijection.
50
Proof. Clearly if two tangent vectors look the same in (U, f ) then they must
look the same in all charts by the rule (5.20), so evf is an injection. Now pick
any vector v Rn . Let (U0 , f0 ) = (U, f ), and define a function : Ax Rn
by:
: (Ui , fi ) 7 fi = Di0 |f0 (x) (v)
If we pick any two charts (U1 , f1 ), (U2 , f2 ) Ax , the transition functions obey
the equation
20 = 21 10
(this makes sense in some neighbourhood of f0 (x)), so by the chain rule:
f2 = D20 |f0 (x)(v) = D21 |f1 (x) D10 |f0 (x) (v) = D21 |f1 (x) (f1 )
So our function obeys the rule (5.20), hence its an element of Tx X. Since
evf () = v this shows that evf is a surjection.
If we have two different charts in Ax then the bijections evf1 and evf2 are
related by:
evf2 = D21 |f1 (x) evf1
(5.22)
The vector space structure on Tx X is much more obvious than the one on
Tx X. For two elements , Tx X, we can define + to be the function
+ : (U, f ) 7 f + f Rn
and this still obeys the rule (5.20), so it is an element of Tx X. Scalar multiplication is similar, and it follows immediately that the function evf is a
linear isomorphism, for any chart (U, f ).
Now we can prove that our two definitions of tangent vectors, Definition
5.5 and Definition 5.19, are equivalent.
Proposition 5.23. There is a canonical linear isomorphism between Tx X
and Tx X.
Proof. If we have a tangent vector [] Tx X in the geometers sense, then
we can get a tangent vector Tx X in the physicists sense by considering
the function:
: Ax Rn
(U, f ) 7 f ()
So there is a natural function:
Tx X Tx X
51
Tx X Rn Tx X
and since both factors are linear isomorphisms, so is their composition.
5.4
Our third definition of tangent vectors is perhaps the most difficult, but in
practice it is the most useful. It takes the point-of-view that a tangent vector
is a direction in which we can take a partial derivative of a function. Again
we start by explaining how this works when our manifold is just an open
subset of Rn .
Let U Rn be an open set. We let C (U ) denote the set of all smooth
functions from U to R, this is an infinite-dimensional vector space under
point-wise addition and scalar multiplication of functions.
Now fix a point x U , and choose a vector v Rn . We have an operation
take the partial derivative at x in the direction v, this is an operator
x,v : C (U ) R
that sends a smooth function h C (U ) to the real number:
x,v (h) = Dh|x (v) R
This operation x,v is a linear map, i.e. we have:
x,v (h1 + h2 ) = x,v (h1 ) + x,v (h2 )
and
(5.25)
n
X
(
yi xi ) x,ei (h) + Hi (
y)
i=1
n
X
x,ei (h))d(
yi xi )
i=1
Rn Derx (U )
53
[]7
Derx (U )
v7x,v
[]7D|0
Rn
Now lets generalize all of this to an arbitrary manifold.
Definition 5.27. Let X be a manifold and let x be a point in X. Let C (X)
denote the vector space of all smooth functions from X to R. A derivation
at x is a linear map
d : C (X) R
obeying the product rule
d(h1 h2 ) = h1 (
x)d(h2 ) + h2 (
x)d(h1 )
for any two functions h1 , h2 C (X). We denote the set of all derivations
at x by Derx (X).
Again its easy to check that Derx (X) is a vector space under the usual
addition and scalar multiplication operations on linear maps.
If is a curve through x, then we can define an operator:
: C (X) R
d(h )
h 7
dt
54
(h) = (h)
since h = h
. So the operator can be described as restrict to
the chart (U, f ), then take the partial derivative along the tangent vector
f () = D
|0 Rn (note that we dont need to check if this is co-ordinate
independent, because we defined in a way that didnt need co-ordinates).
It follows from this description that only depends on the vector D
|0 ,
so it only depends on the equivalence class of in the tangent space Tx X.
So we have a well-defined function:
Tx X Derx X
[] 7
Looking at it in our chart (U, f ), its clear that this function is a linear
injection.
Proposition 5.28. The function [] 7 is a linear isomorphism between
Tx X and Derx (X).
The hard part of this proposition is proving surjectivity, i.e. that every
derivation at x arises in this way. Well need to prove a couple of lemmas
first, and introduce some useful gadgets called bump functions.
You may recall the remarkable function:
:RR
1
e x , for x > 0
x 7
0, for x 0
This function is smooth, it can be differentiated to arbitrary order, because
1
all derivatives of e x tend to zero as x tends to zero. By messing around
with we can create some other nice functions, for example:
(x) =
(x)
(x) + (1 x)
55
Actually, it is not obvious that this function is smooth. In fact it isnt even true
in general! We need a technical condition on our manifold, we must assume that X is
Hausdorff. This is explained in Appendix B, from now on we will assume our manifolds
are Hausdorff without futher comment.
56
So is a function in C (X) which has constant value a in an open neighbourhood W = f 1 (B(0, r)) of x, and constant value b outside a larger open
neighbourhood W 0 = f 1 (B(0, r0 )).
Now we can prove the two lemmas that we need for Proposition 5.28.
Lemma 5.29. Let d be a derivation at x. If h C (X) is identically zero
on some open neighbourhood of x then d(h) = 0.
Proof. Suppose that h is identically zero on a neighbourhood U of x. Let
C (X) be a bump function such that we have neighbourhoods
x W W0 U
with |W 0 and |X\W 0 1. Then h = h, so d(h) = d(h) = 0 by the
product rule.
The subset
{h C (X) ; an open neighbourhood U 3 x with h|U 0}
(5.30)
G
x (X) Gx (U )
given by sending [h] to [h|U ], for any h C (X).
57
G
x (U ) Gx (X)
[
If h is a function on X then (h|
U ) agrees with h on the neighbourhood W .
Also if g is a function on U then (b
g )|U agrees with g on the neighbourhood
W . This proves that the function [g] 7 [b
g ] is the inverse to the function
[h] 7 [h|U ].
The isomorphism in Lemma 5.31 respects multiplication of functions, so
linear maps from G
x (X) to R that obey the product rule are the same thing
as linear maps from G
x (U ) to R that obey the product rule. Using Lemma
5.29, this implies that the map
Derx (U ) Derx (X)
(induced by the map h 7 h|U ) is an isomorphism.
Finally we can prove that Tx X and Derx (X) are the same thing.
Proof of Proposition 5.28. Recall that we have a linear map
Tx X Derx (X)
[] 7
sending a tangent vector to partial derivative operator along that vector. Pick
a chart (U, f ) containing x. We know that Derx (X) is isomorphic to Derx (U ),
and using our co-ordinates f this is isomorphic to Derx (U ). By Lemma 5.26,
the only derivations on U at x are the partial derivative operators x,v . So
the map Tx X Derx (X) is an isomorphism.
58
n X
k
X
i=1
Fj g
vi
xi x yj y
j=1
using the chain rule. So DF |x (x,v ) is the partial derivative operator y,u
where u is the vector we get by multiplying v by the Jacobian matrix.
6
6.1
Vector fields
The tangent bundle
f : Tx X Rn
Associated to our open set U X there is a subset of the tangent bundle
[
T U = 1 (U ) =
Tx X T X
xU
F : T U U Rn
(x, v) 7 f (x), f (v)
f1 : Tx X Rn
f2 : Tx X Rn
and
and the two are related by the derivative of the transition function 21 at
f1 (x):
f2 = D21 |f1 (x) f1
So the first chart gives us a bijection
F1 : T W f1 (W ) Rn
and the second chart gives us a bijection
F2 : T W f2 (W ) Rn
and these two are related by the bijection
21 : f1 (W ) Rn f2 (W ) Rn
(
x, v) 7 21 (
x), D21 |x (v)
This function 21 is smooth, since the entries in the Jacobian matrix D21 |x
depend smoothly on x. The inverse function is just 12 , so this is also
smooth. In particular both 21 and its inverse are continuous, so its a
homeomorphism.
61
f1 : U1 = T 1 \ [0] U1 = (0, 1)
and
f2 : U2 : T 1 \ [ 21 ] U2 = ( 21 , 12 )
So we have an atlas for the tangent bundle T (T 1 ) which has two charts
T U1 = (0, 1) R
and:
T U2 = ( 12 , 12 ) R
The derivative of the transition function 21 at any point is just the identity
map from R to R, so the transition function between our two charts on T (T 1 )
is just:
21 = (21 , 1) : (0, 21 ) t ( 12 , 1) R ( 21 , 0) t (0, 21 ) R
A manifold which has this atlas must be T 1 R.
This example is rather unusual, its not normally true that the tangent
bundle to X is simply X Rn . Manifolds like this are called parallelizable,
well come back to this later on.
Also, this is unfortunately the only example for which its easy to visualise
the tangent bundle, because if X has dimension 2 then T X has dimension 4!
The projection function : T X X is automatically smooth, since if
we choose a chart (U, f ) on X and the corresponding chart (T U, F ) on T X
then the function in these charts becomes the projection map from U Rn
to U . Since this is (the restriction of) a linear map, its derivative at any
point is the same projection map again, which proves that has rank n at
every point and hence is a submersion.
The level sets of are the individual tangent spaces Tx X. Since has no
critical points, these must all be n-dimensional submanifolds of T X. This
fact is also obvious if we look in one of our charts (T U, F ).
Theres another obvious n-dimensional submanifold of T X, given by the
inclusion:
: X 7 T X
x 7 (x, 0)
This function is called the zero section of T X. Looking in a chart (T U, F ),
its obvious that is a smooth immersion, and that the image of is a
submanifold which is diffeomorphic to X.
6.2
such that = 1U . This just says that is a function of the form (1U , ),
so the data of and are exactly the same. However, this second definition
generalizes to an arbitrary manifold.
Definition 6.2. A vector field on X is a function
: X TX
such that = 1X .
So for each point x X, the function selects a vector |x Tx X.
Example 6.3. If X = S 1 , then we saw in Example 5.17 that the tangent
space T(x,y) S 1 to a point (x, y) S 1 can be identified with the subspace of
64
F |U f 1 = (1U , )
65
R T(cos , sin ) S 1
1 7 ( sin , cos )>
So in this chart, the vector field is just a constant function:
1 : U R
This is certainly smooth, which proves that is smooth at every point in S 1
apart from (1, 0), and we can use another polar co-ordinate chart to check
that is smooth at that point too.
From now on well assume that all our vector fields are smooth, unless
we need to specifically state otherwise.
Its easy to find (smooth) vector fields on any manifold X. For example,
pick any chart (U, f ) on X, and choose a bump function on X which
vanishes outside of some open set W U . A vector field on the open set
U is the same thing as a smooth function from U to Rn , and if we choose
one then we can extend it to a (smooth) vector field on X by defining:
inside U
,
=
0, outside U
So the set of all vector fields on X is very large indeed. Also notice that we
can make this set into an (infinite-dimensional) vector space, using point-wise
addition and scalar multiplication.
66
Vector fields are closely related to a nice geometric idea called a flow, as
well now explain.
If X and Y are two manifolds then weve defined a diffeomorphism between X and Y to be a smooth function F : X Y with a smooth inverse
(Definition 4.17). In particular, it makes sense to talk about diffeomorphisms
F :XX
from a manifold to itself. These are the symmetries of a manifold.
Example 6.5. Let X = T 1 = R/Z. For any constant s R, we can define
a bijection
Fs : T 1 T 1
by:
Fs : [t] 7 [t + s]
Its easy to check that Fs is smooth for any s (we can use our atlas on T 1
from Example 2.10), and since the inverse of Fs is Fs this shows that each
Fs is a diffeomorphism.
We checked in Example 4.19 that the function
H : T 1 S1
[t] 7 (cos 2t, sin 2t)
is a diffeomorphism. This implies that for any s the function rotate by 2s
Gs = H Fs H 1 : S 1 S 1
(cos , sin ) 7 cos( + 2s), sin( + 2s)
is a diffeomorphism of S 1 .
In the previous example, we didnt just write down one diffeomorphism,
we wrote down a whole family of them, indexed by the parameter s R,
and in the middle we have the identity function F0 = 1T 1 . Moreover, the
diffeomorphism Fs depends smoothly on s, in the following sense. Put all
of them together to form the function:
F : R T1 T1
(s, [t]) 7 Fs ([t])
The set R T 1 is a 2-dimensional manifold (we wrote down an atlas in
Example 6.1), and its easy to check that this total function F is smooth.
Lets abstract this example.
67
Definition 6.6. Let X be a manifold. A 1-parameter family of diffeomorphisms, or flow, on X is a smooth map
F : (, ) X X
for some positive real number , such that for each s (, ) the function
Fs : X X
x 7 F (s, x)
is a diffeomorphism, and in particular the map F0 is the indentity on X.
It should be obvious that that if take a smooth atlas for X then we can
produce a smooth atlas for (, )X, making it into a manifold of dimension
dim X + 1. So it does makes sense to ask for F to be smooth.
Instead of fixing the parameter s, we can instead choose to fix a point
x X. Then we get a smooth function:
Fx : (, ) X
s 7 F (s, x)
Since Fx (0) = F0 (x) = x, this is a curve through x, so it determines a vector
in the tangent space Tx X. If we do this at all points in x simultaneously
then we produce a vector field on X, which well call F . This vector field
is the infinitesimal version of the flow F , it tells us the direction that every
point will move in if we start to apply the flow.
Lets look at this procedure in co-ordinates. If we pick a chart on X with
codomain U Rn then F will become a smooth function
F = (F1 , ..., Fn ) : (, ) U U
such that F0 = 1U (in fact F might only be defined on an open neighbourhood
F field on U
is:
of the subset {0} U ). The associated vector f
!
1
n
F
F =
=
: U Rn
f
, ... ,
s
s
s
s=0
s=0
s=0
F is a smooth
In particular, its clear that the vector field F is smooth, since f
function.
(here can be any positive real number). The vector field associated to this
flow takes the value
2 sin
G
|(cos , sin ) =
T(cos ,sin ) S 1 R2
2 cos
at the point (cos , sin ) S 1 . Up to the overall scale factor of 2, this is
the vector field that we saw in Example 6.3.
Its in interesting question to ask whether this process can be reversed: if
we have a vector field on X, can we construct a flow on X whose associated
vector field is ?
This is a question about constructing solutions to partial differential equations, and dealing with it properly requires more analysis than we wish to
introduce here. However, the answer is yes, provided that we assume that
X is compact. If we fix a small neighbourhood U X then we can always
construct a flow
F : (, ) U X
whose infinitesimal version is |U , for some value of . If X is non-compact
then these values for might not be bounded above zero over the whole of
X, meaning that we cannot find a global flow F for any positive value of .
However if X is compact then there must some minimal 0 > 0, and we have
a global flow with = 0 .
6.3
Well now see what vector fields look like if we adopt the viewpoints of our
second two definitions of tangent vectors, Definition 5.19 and Definition 5.32.
If is a vector field on X, then for any chart (U, f ) we have a smooth
function : U Rn . If we have two charts (U1 , f1 ) and (U2 , f2 ), then the
two functions
1 : U1 Rn
and
2 : U2 Rn
are related by the derivatives of the transition function 21 . For any point
x U1 U2 , we must have the transformation law:
2 |f2 (x) = D21 |f1 (x) 1 |f1 (x)
In the style of Definition 5.19, we can take these properties to be the
definition of a vector field.
69
U2 = R
and
21 : R \ 0 R \ 0
1
x 7
x
You can think of this data as alternative definition of the manifold S 1 ; it says
we take two copies of R and glue them together using 21 .
70
2 : R R
and
such that
x)
2 ( x1 ) = x12 1 (
for all x 6= 0. Its very easy to construct vector fields on S 1 using this
definition, just take any smooth function 1 : R R and define:
2 : R \ 0 R
x 7
x2 1 ( 1 )
x
1 : x 7 x
2 : x 7
x
1 : x 7 x2
2 : x 7 1
n
X
i=1
71
h
i
xi
Since and h are both smooth functions, the function (h)
is also smooth.
This means we can view our vector field as an operator:
: C (U ) C (U )
h 7 (h)
When we want to think in this way its common to write vector fields in the
form:
n
X
=
i
xi
i=1
Recall that every operator x,| x is a derivation at x, i.e. it satisfies the
product rule (5.25). This implies that the operator satisfies the following
version of the product rule:
1 h2 ) = h1 (h
2 ) + h2 (h
1)
(h
(6.10)
This operator is local, meaning that the restriction of the function (h) to
some open subset U X only depends on h|U , it doesnt depend on the
behaviour of h outside U . This means we can pick a chart (U, f ) on X and
translate everything over to the open set U Rn . The vector field becomes
a smooth function : U Rn , and the function (h), looked at in these co h)
as we defined before. This shows that
ordinates, is just the function (
(h) is a smooth function, so is an operator:
: C (X) C (X)
Furthermore the product rule (6.10) holds, since we can check it by working
in co-ordinate charts.
Definition 6.11. A derivation on a manifold X is a linear map
D : C (X) C (X)
such that the product rule
D(h1 h2 ) = h1 D(h2 ) + h2 D(h1 )
holds for all h1 , h2 C (X). The set of all derivations on X is denoted by
Der(X).
Weve seen that any smooth vector field defines a derivation in Der(X).
The converse is also true:
Proposition 6.12. Any derivation D Der(X) defines a smooth vector
field.
Proof. Pick a D Der(X). For a fixed point x X, we can define a linear
operator
D|x : C (X) R
by:
D|x : h 7 D(h) |x
The product rule (6.10) implies that D|x is a derivation at x, so by Proposition 5.28 it must be the partial derivative operator associated to some tangent
vector in Tx X. Hence the function : x 7 D|x is a vector field on X.
It remains to show that is smooth. Pick any chart (U, f ), and let
x1 , ..., xn be the co-ordinate functions on U . In these co-ordinates, we can
write the vector field as
n
X
i
xi
i=1
73
for some functions 1 , ..., n : U R, and what we need to show is that each
of these functions i is smooth. More specifically, lets fix a point y U , set
y = f (y) U , and prove that each i is smooth at y.
Let e be a bump function on U which is constantly equal to 1 inside
of y, and constantly equal to zero outside some
some neighbourhood W
larger neighbourhood. Then we can use this bump function e to extend each
co-ordinate function xi C (U ) to a smooth function i C (X), by
defining
e f inside U
(
xi )
i =
0 outside U
(this is the same trick we used to prove Lemma 5.31). Then if we write i in
the chart (U, f ), we get a function i C (U ) which agrees with xi inside
of y.
the open neighbourhood W
By definition, applying our derivation D must send each i to a smooth
function D(i ) C (X). Lets compute this function D(i ) inside the open
) of y. Its value at any point x W is given by
neighbourhood W = f 1 (W
applying the operator D|x to i , and we can compute this in the chart (U, f )
and see that
n
X
i
= i |f (x)
j |f (x)
D(i )|x =
j
f
(x)
j=1
. Since D(i ) is smooth, i must be smooth
since i xi in the open set W
, and in particular smooth at y.
inside W
So the set of all vector fields on X is exactly Der(X). Its easy to check
that Der(X) is a vector space under the usual addition and scalar multipliation of linear maps, and its clear that this vector space structure is the same
as the one we saw in our previous definitions.
This characterization of vector fields reveals an additional interesting
structure, the Lie bracket.
Suppose we have two derivations D, E Der(X). The composition D E
is a linear map from C (X) to C (X), but it might not be a derivation.
However, it is easy to compute that the commutator of D and E
[D, E] : C (X) C (X)
h 7 D E(h) E D(h)
does obey the product rule, and hence is also an element of Der(X).
74
n
X
i=1
i
xi
and
n
X
i=1
i
xi
=
,
j
j
xj
xj xi
i=1 j=1
The Lie bracket also has a geometric interpretation in terms of flows,
which were going to quickly sketch. Firstly, suppose that we have a single
diffeomorphism:
F :XX
Then for any point x X, we have an isomorphism:
DF |x : Tx X TF (x) X
Now suppose we also have a vector field on X. We can define a new vector
field F by pulling-back along F , meaning that we define:
(F )|x = (DF |x )1 |F (x)
Notice that this trick only works if F is a diffeomorphism, it wont work for
more general smooth functions.
Now suppose we have a flow
F : (, ) X X
which we recall means a smooth family of diffeomorphisms Fs : X X, with
F0 = 1X . Then we can use F to turn our vector field into a 1-parameter
75
family of vector fields, since for every s (, ) we have a vector field Fs
on X. For s = 0 we get the original vector field F0 () = . If we fix a single
point x X then we have 1-parameter family of vectors in the tangent
space Tx X, given by
(Fs )|x Tx X
for each s (, ). The derivative of this family at s = 0 is a new vector
(Fs )|x
Tx X
s s=0
so if we evaluate this at all points x X then we obtain a new vector field.
This new vector field measures the infinitesimal change in the vector field
when we apply the flow F .
Now the flow F has its own vector field F associated to it, which is the
infinitesimal version of F . We claim that the infinitesimal change in when
we apply the flow F is given by the vector field:
[ F , ]
Well give a sketch proof of this claim. The statement can be checked in
co-ordinates, so let
F : (, ) U U
be a flow and let
: U Rn
be a vector field. For any point x U , we have
x + O(s2 )
F (s, x) = x + s|
where is the infinitesimal version of F . Hence the Jacobian matrix of F at
(s, x) is
DF |(s,x) = I + sJ + O(s2 )
where I is the identity matrix and J is the matrix whose entries are:
i
Jij =
xj
x
76
On the other hand, the value of the vector field at the point F (s, x) is
+ sK |
x + O(s2 )
|
F (s,
x) = |x
where K is the matrix whose entries are:
i
Kij =
xj
So applying our flow F to the vector field gives a family of vector fields
Fs whose values are:
1
+ sK |
x sJ |
x + O(s2 )
(Fs )|x = DF |(s,x)
|
F (s,
x) = |x
If we take the partial derivative with respect to s, and then set s to zero, we
get the vector field
x J |
x = ,
K |
x
as claimed.
6.4
Vector bundles
Weve seen that for every manifold X theres an associated tangent bundle,
which is a smooth manifold T X coming with a smooth surjection
: TX X
such that every level set 1 (x) = Tx X is a vector space. This is a very rich
mathematical structure, its an example of something called a vector bundle.
Let X be a manifold of dimension n. Informally, a vector bundle over a
manifold X is a collection of vector spaces {Ex }, indexed by the points of
x. These vector spaces have to fit together to give a smooth manifold E,
equipped with a smooth map : E X whose level set over x X is the
associated vector space 1 (x) = Ex . For example, for any manifold X, and
any integer r, there is a vector bundle
: E = X Rr X
where is the projection map : (v, x) 7 x. Its easy to show that theres
a smooth structure on E making it into a manifold (of dimension n + r), and
that is smooth. Obviously the level set of at any point x X is the
vector space Rr . This is called the trivial vector bundle of rank r.
77
The tangent bundle is not usually of this form, in general we cant canonically identify Tx X with Rn so its not usually true that T X = X Rn .
However if we pick a chart (U, f ), then within the open set U X it is true
that T U = U Rn , since our co-ordinates give us this bijection. So within
small neighbourhoods in X, the tangent bundle looks like the trivial bundle
of rank n. This condition, of being locally trivial, is the defining property
of a vector bundle.
Definition 6.14. Let X be a manifold of dimension n, let E be a manifold
of dimension n + r, and let
:EX
be a smooth surjection. For each x X we denote the level set of by
Ex = 1 (x).
Suppose that for every point x X we have specified the structure of an
r-dimensional vector space on Ex . We call this structure a vector bundle if
there exist atlases {(Ui , fi ), i I} for X and {(Vi , gi ), i I} for E (indexed
by the same set I) with the following properties:
Vi = 1 (Ui ), for each i I.
Vi = Ui Rr Rn+r , for each i I.
For each i I, the square
Vi
gi
Ui Rr
Ui
fi
Ui
78
F : E1 E2
such that 2 F = 1 , and such that the induced function
Fx : (E1 )x (E2 )x
is a linear isomorphism, for each x X.
So an isomorphism of vector bundles is a bijection that preserves all the
structure of a vector bundle. In particular if two vector bundles over X are
isomorphic they must obviously have the same rank.
Definition 6.18. A rank r vector bundle : E X is called trivial if it is
isomorphic to the trivial vector bundle X Rr .
Here is one way to tell if a vector bundle is trivial:
Proposition 6.19. Let : E X be a vector bundle of rank r. Then E
is trivial iff there exist r sections 1 , ..., r of E such that, for every point
x E, the vectors
1 |x , ..., r |x Ex
form a basis of Ex .
Proof. If E is the trivial vector bundle X Rr then we can just pick any
basis e1 , ..., er for Rr and consider the constant sections
i : x 7 ei for each
i, which are evidently smooth. More generally if F : X Rr E is an
isomorphism of vector bundles then we can define r sections of E by:
i = F
i : x 7 Fx (ei ) Ex
These are smooth since both F and
i are smooth, and give a basis of Ex
since Fx is an isomorphism of vector spaces.
80
x,
r
X
!
vi i |x
i=1
Obviously F commutes with the projection maps, and for each x X the
map Fx is linear and sends the standard basis of Rr to the basis 1 |x , ..., r |x
of Ex . Hence each Fx is an isomorphism of vector spaces, and it follows that
F is a bijection.
Now pick a chart (U, f ) on X and a chart (V, g) on E of the form specified
in Definition 6.14, and choose the corresponding chart U Rr on X Rr . In
these charts, each section i is a smooth function
i = (
1i , ....,
ri ) : U Rr
and F is the function
F : U Rr U Rr
given by the smooth family of invertible r-by-r matrices Mx whose entries
are
ji |x . The inverse function F 1 is given by the family of matrices Mx1 ,
whose entries will also vary smoothly with x since they are rational functions
of the entries in Mx . Hence both F and F 1 are smooth, so we have shown
that F is an isomorphism of vector bundles.
Definition 6.20. A manifold X is said to be parallelizable iff its tangent
bundle T X is trivial.
Example 6.21. Let X = S 1 . In Example 6.3 we found a smooth vector field
on S 1 which was not equal to zero at any point. Since S 1 is 1-dimensional,
this means that gives a basis of the tangent space at every point. So by
Proposition 6.19 the bundle T S 1 is trivial, and S 1 is parallelizable.
The manfold S 2 is not parallelizable, because of the following fact:
Theorem 6.22 (Hairy ball theorem). Any vector field on S 2 must be equal
to zero at some point.
Consequently it is impossible to find a pair of vector fields 1 , 2 on S 2
that form a basis of the tangent space at every point.
Theorem 6.22 is is a very nice result. It implies for example that at any
moment in time there must be a point on the Earth where the wind speed
81
is zero, and also that you cannot groom a spherical dog without leaving a
protruding tuft of hair at one point. The proof is not very difficult, but
unfortunately it requires some algebraic topology that doesnt form a part of
this course.
Theorem 6.22 is true for any even-dimensional sphere S 2n , so no evendimensional sphere is parallelizable. In fact the only parallelizable spheres
are S 1 , S 3 and S 7 , but this is rather harder to prove.
7
7.1
In this section were going to look at the dual objects to tangent vectors,
which are called covectors. Well begin by recalling some basic facts about
dual vector spaces.
Let V be a vector space, of dimension n. Recall that the dual vector space
to V is the space
V ? = Hom(V, R)
of all linear maps from V to R. The dimension of V is also n, if we pick a
basis {e1 , ..., en } for V there is a corresponding dual basis {1 , ..., n }, where
i V ? is the linear map defined by:
i : V R
1, j = i
ej 7
0, i =
6 j
If V is just Rn , the space of length-n column vectors, then the dual space V ? is
the space of length-n row vectors, which means that theres a canonical way to
identify (Rn )? with Rn . Under this identification, the standard basis becomes
its own dual basis, and the process of evaluating a row vector u (Rn )? on
a column vector v Rn becomes the dot product of vectors.
If W is a second vector space (of dimension m) and
F :V W
is a linear map, then there is a corresponding dual linear map
F? : W? V ?
which sends a vector u W ? to a vector F ? (u) V ? defined by:
F ? (u) : V R
v 7 u(F (v))
82
If we compose two linear maps F and G then its easy to see that the dual
of the composed map is:
(G F )? = F ? G?
If we pick a basis for both V and W then the linear map F can be expressed
as an m-by-n matrix:
F : Rn Rm
The dual map F ? can be an expressed as an n-by-m matrix, using the corresponding dual bases of V ? and W ? , and its easy to calculate that it becomes
the transpose matrix:
F > : Rm Rn
Now we return to manifolds.
Definition 7.1. Let x be a point in a manifold X. The cotangent space
to X at x is the dual vector space to Tx X. Elements of the cotangent space
are called covectors.
Its conventional to denote the cotangent space to X at x by
Tx? X
(rather than (Tx X)? ). We know that if we pick a chart (U, f ) containing x
then we get a linear isomorphism:
f : Tx X Rn
The dual linear map to f gives a linear isomorphism
?f : Rn Tx? X
which we can invert to give:
(?f )1 : Tx? X Rn
So if we work in co-ordinates then we can identify both the tangent space
and the cotangent space to x with the vector space Rn . Its only when we
change co-ordinates that we see the difference between them.
Recall that if we have two charts (U1 , f1 ) and (U2 , f2 ) containing x then
tangent vectors change according to the transformation law
f2 = D21 |f1 (x) f1
83
where 21 is the transition function between the two charts. We can also
write this as
1
f1 = D21 |f1 (x)
f2 = D12 |f2 (x) f2
using the inverse transition function 12 . Taking the dual of this equation
tells us that
?
?f1 = ?f2 D12 |f2 (x)
and then inverting gives
(?f1 )1 =
? 1
so:
(f2 )1 = D12 |f2 (x)
?
(?f2 )1
(f1 )1
(7.2)
Proof. This is proved in exactly the same way as Proposition 5.23. Let
Tx? X denote the set of all such functions u, this has an obvious vector space
structure. For any chart (U, f ), the evaluation map
f : Tx? X Rn
u 7 uf
is a linear isomorphism, by exactly the same argument that proved Lemma
5.21. Then the evident function from Tx? X to Tx? X must be a linear isomorphism, since if we pick any chart then we can factor it as:
Tx? X
(?f )1
f1
R Tx? X
The analogue for covectors of the algebraists definition of tangent vectors (Definition 5.32) is less obvious. It will take us a little while to describe,
but in the end it is quite simple.
Recall that we may think of a tangent vector v Tx X as a partial
derivative operator
v : C (X) R
which is an element of Derx (X). This operator can be evaluated either by
differentiating functions along some curve such that [] = v Tx X, or by
choosing co-ordinates and then performing partial differentiation along the
corresponding vector in Rn .
Now lets turn this definition around. If we fix a smooth function h
C (X) then we can produce a linear map from Tx X to R, which we will
denote by dh|x , by defining:
dh|x : Tx X R
v 7 v (h)
So to every smooth function h C (X) there is an associated covector
dh|x Tx? X.
: R2 R be the function h
: (x, y) 7 x2 + y. The
Example 7.4. Let h
2
tangent space to a point (x, y) R is spanned by the two partial derivative
to 2x and
to 1.
85
h
y : y,e 7 y,e (h)
=
dh|
i
i
xi
y
h
y =
dh|
x1 |y + ... +
xn |y Tx? U
(7.5)
d
d
x1
xn
y
>
h
h
, ... ,
Rn
x1
xn
f (x)
f (x)
f (x) )>
We know that dh|x is a covector, so we know that the column vector (Dh|
must obey the transformation law (7.2), but for the sake of completeness lets
verify this explicitly. In fact weve already done the calculation, because we
know that when we change charts the Jacobian matrix changes by the rule
(4.8), so
2 |f (x) = Dh
1 |f (x) D12 |f (x)
Dh
2
1
2
and transposing this gives
2 |f (x)
Dh
2
>
>
>
1 |f (x)
Dh
1
as required.
Now we give the algebraists definition of a covector. As weve just seen,
to any function h C (X) theres an associated covector dh|x . Furthermore,
its easy to check that this function
C (X) Tx? X
h 7 dh|x
is itself a linear map. However, the vector space Tx? X is finite-dimensional,
whereas the vector space C (X) is (uncountably) infinite-dimensional, so
this linear map must have an infinite-dimensional kernel. Using our expression for dh|x in co-ordinates, we can see that dh|x = 0 iff in any chart (U, f )
containing x, the corresponding Jacobian matrix
f (x) : Rn R
Dh|
is the zero row vector. This is precisely the statement that the function h
has rank zero at the point x, because the rank of h at x must be either one
f (x) 6= 0 then
or zero (since it is a smooth function from X to R), and if Dh|
?
it has rank one. So the kernel of our linear map C (X) Tx X is the subset
Rx (X) C (X)
of functions which have rank zero at x. In particular Rx (X) must be a
subspace.
Proposition 7.6. We have an isomorphism
87
Proof. Weve seen that h 7 dh|x defines a linear map from C (X) to Tx? X,
and that the kernel of this map is Rx (X). It only remains to show that
this map is a surjection, since then the Proposition follows by the First
Isomorphism Theorem.
Pick a chart (U, f ) containing x, and let y = f (x). Then we can identify
?
xi |y, ..., d
xn |y dual to the standard basis
Tx X with Ty? U , and this has a basis d
y,e1 , ..., y,en of Dery(U ). Its enough to show that each basis vector d
xi |y
can be obtained from some smooth function on X.
Using a bump function (as in the proof of Proposition 6.12), for each i
we can find a smooth function xi C (X) such that when we write xi in
our chart, it agrees with the co-ordinate function xi in some neighbourhood
of y. This means that if we evaluate the covector dxi |x against any tangent
vector y,v Dery(U ) we get the answer:
y,v (
xi ) = d
xi |y (y,v )
So in this chart, dxi |x is exactly d
x|i .
So from this point-of-view, a covector at x is just an equivalence class of
smooth functions on X, where we regard two functions as equivalent if their
difference has rank zero at x.
Incidentally, this shows yet another possible way to define the tangent
space Tx X. We could define Tx? X to be the quotient of C (X) by Rx (X),
and then define Tx X to be the dual space to Tx? X. This is certainly the
quickest definition to write down!
7.2
Just as we did for the tangent bundle, we can take all the cotangent spaces
Tx X for each x X, and assemble them together to get a set
[
T ?X =
Tx? X
xX
T ? U U Rn
(x, u) 7 f (x), (?f )1 (u)
88
f1 (W ) Rn f2 (W ) Rn
(
x, u) 7 21 (
x), (D12 |f2 (x) )> (
u)
This means we can put a topology on T ? X such that these bijections are
co-ordinate charts, then we have a smooth atlas of exactly the form required
by Definition 6.14.
Definition 7.7. Let X be a manifold. A covector field, or one-form, on
X is a section of the cotangent bundle, i.e. a smooth function
: X T ?X
such that = 1X .
The reason for the name one-form is that this is the first case of a more
general object called a p-form, where p can be any natural number. We will
meet p-forms later on.
If our manifold is simply an open set U Rn then the tangent bundle is
just U Rn and a one-form is the same thing as a smooth function:
: U Rn
So on U , its difficult to see the difference between one-forms and vector
fields. On more complicated manifolds we do see the difference, because the
way that one-forms change when we change co-ordinates is different from the
way that vector fields change. If is a one-form on X, and we have two
charts (U1 , f1 ) and (U2 , f2 ), then in each chart becomes a smooth function:
1 : U1 Rn
2 : U2 Rn
For points x U1 U2 , we have the transformation law
(7.8)
and this is different from the transformation law for vector fields.
As we did for vector fields (Proposition 6.8), we could take these properties as an alternative definition of a one-form. Also note that we can specify
a one-form on X by choosing an atlas A = {(Ui , fi )} for X and choosing
functions
i : Ui Rn
89
>
h
h
y 7
, ... ,
x1
xn
y
vary smoothly
This is a smooth function, since the partial derivatives of h
to be one of the co-ordinate
with y. In particular, notice that that we set h
functions xi on U , then d
xi is the constant function sending every point in
h
h
d
x1 + ... +
d
xn
x1
xn
f 1 : (, ) S 1 \ (1, 0)
7 (cos , sin )
90
By definition, F ? d
yj is the function
F ? d
yj : U Rn
which sends a point x to the column vector
!>
Fj
Fj
Rn
, ... ,
x1
x1
x
that we get by applying the matrix (DF |x )> to the j-th standard basis vector
in Rk . This is a smooth function, since the partial derivatives of Fj vary
smoothly with x. Also, we have that
F ? d
yj =
Fj
Fj
d
x1 + ... +
d
xn
x1
xn
(7.10)
=
1 d
y1 + ... +
k d
yk
for k smooth functions
1 , ...,
k C (V ). Then
yk |F (x)
k |F (x) d
y1 |F (x) + ... +
F ?
|x = (DF |x )?
1 |F (x) d
so
F ?
= (
1 F )F ? d
y1 + ... + (
k F )F ? d
yk
since each (DF |x )? is a linear map. Each F ? d
yk is a smooth function from U
n
to R , and each
k F is a smooth function from U to R, so F ?
is smooth.
This local calculation proves that on general manifolds, the pull-back of
a one-form is indeed a (smooth) one-form.
Example 7.11. Consider the inclusion map:
: S 1 R2
On R2 we have two (constant) one-forms dx and dy. Lets calculate the
one-forms dx and ? dy on S 1 . If we think of the tangent space to a point
(x, y) S 1 as the subspace
T(x,y) S 1 R2
spanned by (y, x)> , then D|(x,y) is just the inclusion of this subspace, and
hence
? dx|(x,y) : T(x,y) S 1 R
92
is the linear map which sends (y, x) to y. Similarly dy|(x,y) sends (y, x)
to x.
Now lets look at these one-forms on S 1 in co-ordinates. Weve seen
(Examples 5.17 and 6.4) that if use polar co-ordinates (x, y) = (cos , sin )
then the tangent vector (y, x) corresponds to the unit tangent vector
.
This means that in this chart, dx becomes the one-form sin d and dy
becomes the one-form cos d.
If we choose our one-form on Y to be of the form = dh for some
h C (Y ), then there is another way to say what the pull-back of is.
Proposition 7.12. If F : X Y is a smooth function and h C (Y )
then:
F ? dh = d(h F )
Proof. Fix a point x X and let y = F (x). By definition, (F ? dh)|x is the
covector (DF |x )? (dh|y ) Tx? X. Now choose any d Derx (X) = Tx X. By
the definition of a dual linear map, we have that (DF |x )? (dh|y ) sends d to
the real number dh|y (DF |x (d)). By the definition of the covector dh|y this
number is (DF |x (d))(h), and this equals d(h F ) by Lemma 5.33. So weve
shown, for any point x X, that (F ? dh)|x is the covector:
(F ? dh)|x : Tx X R
d 7 d(h F )
This proves that F ? dh = d(h F ).
In particular, suppose that U Rn and V Rk are open sets, and
F = (F1 , .., Fk ) : U V is a smooth function. If we take one of the coordinate functions yj C (V ), then F yj is just Fj , and the proposition
weve just proved says that
F ? d
yj = dFj
which is exactly what we saw in the equation (7.10).
C (R2 )
Example 7.13. Let : S 1 , R2 be the inclusion map, and let h
2
1
: (x, y) 7 x . Then h = h
C (S ) is the function
be the function h
considered in Example 7.9, and we calculated there that if we look at dh in
polar co-ordinates we get 2 sin cos d. On the other hand, we have that:
= 2x dx
dh
In Example 7.11 we saw that if we look at dx in in polar co-ordinates we will
in polar co-ordinates we must
get sin d, and therefore if we look at dh
get 2 cos sin d. These two calculations had to give the same answer,
Its worth noticing that the transformation law for one-forms (7.8) is
actually a special case of this pull-back operation. Suppose we have two
charts (U1 , f1 ) and (U2 , f2 ) on X, so we have a transition function:
21 : f1 (U1 U2 ) f2 (U1 U2 )
Now pick a one-form on X, which in our charts becomes:
1 : U1 Rn
and
2 : U2 Rn
2 = ?12
1
on the open set f2 (U1 U2 ).
Differential forms
8.1
2 V ?
Now suppose we have two elements u, u of the space V ? . We can combine
them to form an element of 2 V , by setting:
u u : V V 7 V
(v, v) 7 u(v)
u(
v ) u(
v )
u(v)
Clearly u u is an antisymmetric bilinear map. We call it the wedge product
of u and u. This wedge product is an important structure, we can think of
it as a kind of multiplication:
: V ? V ? 2 V ?
(u, u) 7 u u
Its a straight-forward exercise to check that this product is itself bilinear.
Its also antisymmetric, since its clear from the definition that:
u u =
uu
Now pick a basis e1 , ..., en for V , and let 1 , ..., n be the dual basis for V ? .
Choose a pair i, j [1, n] with i < j, and form the bilinear map i j 2 V ? .
Applying this to pairs of basis vectors in V we get:
1, s = i and t = j
1, s = j and t = i
i j : (es , et ) 7
0,
otherwise
So i j corresponds to the matrix with a 1 in the (i, j) position (which is
above the diagonal), a 1 in the (j, i) position (which is below the diagonal),
and zeroes everywhere else. Clearly this set of matrices forms a basis for the
space of all antisymmetric n-by-n matrices, so the set
{i j ; i < j} 2 V ?
95
n
=
2
We can use these bases to describe the wedge product explicitly. If we take
two elements
u = 1 1 + ... + n n
and
u = 1 1 + ... + n n
V?
(here 1 , ..., n and 1 , ..., n are just real numbers) then their wedge product
is:
u u = 1 2 2 1 1 2 + 1 3 3 1 1 3 + ...
... + n1 n n n1 n1 n
Example 8.1. Let V = R3 , and e1 , e2 , e3 be the standard basis. Then V ? is
also
R3 , and 1 , 2 , 3 is again the standard basis. The dimension of 2 V ? is
3
= 3, and it has a basis:
2
{2 3 , 1 3 , 1 2 }
Using these bases, the wedge product of two vectors (1 , 2 , 3 ) and (1 , 2 , 3 )
is:
(2 3 3 2 , 1 3 3 1 , 1 2 2 1 )
If we flip the sign of the basis vector 1 3 , then the formula above becomes
the usual cross-product of vectors in R3 . This explains why there is no
direct
analogue of the cross-product in other dimensions, since if n 6= 3 then
n
6= n. In other dimensions, the cross-product of two vectors is really the
2
n
wedge product, and it lands in R( 2 ) .
We know that a linear map F : V W induces a dual linear map
F : W ? V ? . It also induces a linear map
?
2 F ? : 2 W ? 2 V ?
defined by
k
2
96
2 (G F )? = 2 F ? 2 G?
In particular, if F is an isomorphism, then so is 2 F ? . If you know what
a functor is, this says that the operation which sends V to 2 V ? is a contravariant functor (as is the operation which sends V to V ? ).
2 V ? to
c : V p R
where V p means V ...(p times)... V . If we choose a basis e1 , ..., en for V
then c gives us a p-dimensional array of numbers
Ci1 ,...,ip = c(ei1 , ..., eip )
by evaluating c on each p-tuple of basis vectors, and conversely any such
array of numbers determines a p-linear map by extending linearly in each
argument. Hopefully its clear that the set of all p-linear maps from V to R
is a vector space, of dimension np .
We say that c is antisymmetric if c flips sign when we swap any two of
its arguments, i.e.
c(v1 , ..., vp ) = c(v(1) , ..., v(p) )
for any transposition acting on the set {1, ..., p}. In particular this means
that if we set any two of its arguments to be the same vector, then c must
give the answer zero. If we apply a more general permutation Sp , we
must have:
c(v1 , ..., vp ) = (1) c(v(1) , ..., v(p) )
where (1) is our notation for the sign of the permutation .
The antisymmetric maps form a subspace of the space of all p-linear maps
from V to R, and we denote this vector space by:
p V ?
Weve seen that two elements of V ? can be wedged together to get an
element of 2 V ? . Similarly, if we have p elements u1 , ..., up of V ? , then we
97
(8.2)
Sp
This map is clearly linear in each argument, and by construction its antisymmetric, so it is indeed an element of p V ? .
Hence weve defined a p-fold wedge product:
(V ? )p p V ?
(u1 , ..., up ) 7 u1 ... up
This product is p-linear and antisymmetric, since the expression (8.2) is linear
in each ui , and changes sign if we switch any ui and uj . In particular if any
two ui and uj are equal then we get the zero element of p V ? .
Now suppose we choose a basis e1 , ..., en for V , so we get a dual basis
1 , ..., n for V ? . We can produce elements in p V ? by picking an p-tuple
(i1 , ..., ip ) of integers in [1, n] and then forming the wedge product i1 ...ip .
If any entries in our p-tuple are repeated then this product must be zero.
If our p-tuple contains no repeated entries, then it must be of the form
((j1 ), ..., (jp )) for some correctly-ordered p-tuple j1 < ... < jp and some
permutation Sp . Then antisymmetry implies that:
(j1 ) ... (jp ) = (1) j1 ... jp
So up to sign, this procedure creates one element of
[1, n] of size p.
Proposition 8.3. Let e1 , ..., en be a basis for V , and let 1 , ..., n be the dual
basis for V ? . Then the set of elements
i1 i2 .... ip | 1 i1 < i2 < ... < ip n p V ?
is a basis. In particular:
dim
n
V =
p
?
If V = Rn and e1 , ..., en is the standard basis, then this proposition provides us with a basis for the space p (Rn )? . This means that we can identify
n
p (Rn )? with R(p) if we wish, but this is not quite canonical, since theres
no preferred way to order the basis vectors in p (Rn )? .
98
Proof. Choose a correctly-ordered p-tuple i1 < ... < ip with each entry in
[1, n], and form the p-linear map i1 ... ip p V ? . Now take an arbitrary
p-tuple (j1 , ..., jp ) of numbers from the set [1, n], and consider evaluating the
map i1 ... ip on the p-tuple of basis vectors
(ej1 , ..., ejp ) V p
using the defining formula (8.2). We can only get a non-zero result if the ptuple (j1 , ..., jp ) is a permutation of the p-tuple (i1 , ..., ip ), in particular there
must be no repetitions in the first p-tuple. So we have
i1 ... ip : (ei(1) , ..., ei(p) ) 7 (1)
for each Sp , and it vanishes on every other p-tuple of basis vectors.
A general antisymmetric p-linear map c is determined by its values on
each p-tuple of basis vectors for V . It must vanish on p-tuples containing
any repetition, and if j1 < ... < jp is a correctly-ordered p-tuple then we
must have
c : (ej(1) , ..., ej(p) ) 7 (1) c(ej1 , ..., ejp )
for each S p . This means that c can be written as a linear combination
X
c=
c(ei1 , ..., eip )i1 ... ip
i1 <...<ip
p F ? : p W ? p V ?
just as we did in the case p = 2, by defining:
Sp
= det(M )
so:
100
(8.5)
(8.7)
This product is zero if the two subsets {i1 , ..., ip } and {j1 , ..., jq } [1, n] are
not disjoint. If they are disjoint, theres some shuffle permutation Sp+q
that returns the (p + q)-tuple (i1 , ..., iq , j1 , ..., jq ) to its correct order, and
the product (8.7) is equal to (1) times the corresponding basis vector in
p+q V ? .
This defines the wedge product (8.6) on each pair of basis vectors, then we
can extend bilinearly. However, we need to check that this definition doesnt
depend on our choice of basis, and this is guaranteed by the next lemma.
Lemma 8.8. For each p, q there is a unique bilinear map from p V ? q V ?
to p+q V ? which makes the following triangle commute:
p V ? q V ?
(V ? )p (V ? )q
p+q V ?
What this lemma says is that if we have two decomposable elements
u1 ... up p V ?
and
u1 ... uq q V ?
Proof of Lemma 8.8. Suppose weve found such a bilinear map (the vertical
arrow in the triangle). Choose a basis 1 , ..., n of V ? . If we take any two basis
vectors i1 ... ip p V ? and j1 ... jq q V ? then commutativity
of the triangle forces the product of these two basis vectors to be given by
the expression (8.7), and then bilinearity determines all other products. This
proves uniqueness.
For existence, we have to check that this product weve defined using our
basis really does make the triangle commute, and by multi-linearity its sufficient to check this on any (p + q)-tuple (i1 , ..., ip , j1 , .., jq ) of basis vectors
for V ? . If any entries are repeated in this (p + q)-tuple then going either way
around the triangle gives the answer zero. If no entries are repeated then
going either way around the triangle gives the corresponding basis vector
in p+q V ? , so we just need to check that the signs match. If go diagonally
across the triangle then we get the sign of the permutation Sp+q that restores this (p+q)-tuple to its correct order. We can factor as first correctly
order (i1 , ..., ip ), then correctly order (j1 , ..., jq ), then shuffle them together,
and this corresponds exactly to the sign that we pick up by going the other
way around the triangle.
This extended version of the wedge product behaves very nicely, as the
next proposition shows.
Proposition 8.9.
have:
p V ? , c q V ? and c r V ? , we
c (
c c) = (c c) c p+q+r V ?
(ii) For any c p V ? and c q V ? we have:
c c = (1)pq c c p+q V ?
(iii) If we have a linear map F : U V then for any c
c q V ? we have:
p+q F ? (c c) = p F ? (c) q F ? (c)
p V ? and
c = u1 ... uq ,
c = u1 ... ur
Property (i) is obvious. Property (ii) just says that the sign of the permutation (p+1)...(p+q)1....p Sp+q is (1)pq . Property (iii) follows immediately
from the observation (8.4).
102
We can also extend the wedge product down to the case when p = 0 (or
q = 0), by declaring that if 0 V ? = R and c q V ? then c is just c,
the scalar multiple. Its trivial to check that the properties in Proposition
8.9 continue to hold in this case.
If we take the direct sum of all our
V ? =
n
X
p V ?
i=0
called the exterior algebra of V . We can give V ? a (bilinear) multiplication by using our wedge product for each component. Property (i) in
Proposition 8.9 says that this structure is an associative algebra, and it has
a unit 1 0 R. Property (ii) in the proposition says that this algebra is
supercommutative. Property (iii) says that is a contravariant functor from
vector spaces to algebras.
If these words are unfamiliar, dont worry, you may safely ignore the last
paragraph!
?
8.2
p-forms
p Tx? X
xX
which is called the p-th wedge power ofthe cotangent bundle. This set is in
fact a vector bundle over X, of rank np . To see this, recall that choosing a
p 1
: p Tx? X p (Rn )?
f
n
We can further identify p (Rn )? with R( p ) , once weve chosen an ordering
of the standard basis of p (Rn )? provided by Proposition 8.3. If we do this
over the whole chart, we get a bijection:
n
p T ? U
U R( p )
103
Then we can proceed by exactly the same argument that we used to show
that T X and T ? X were vector bundles.
Definition 8.10. A p-form on X is a section of the vector bundle
p T ? X.
: U p (Rn )?
Now recall that the co-ordinate functions x1 , ..., xn on U give us a set of
one-forms on U
d
x1 , ..., d
xn
and if we evaluate this set of one-forms at any point y then we get the
standard basis for the cotangent space Ty? U
= Rn . This means that if we
take a correctly-ordered p-tuple i1 < ... < ip , and form the function
d
xi1 ... d
xip : U p (Rn )?
then at any point y U this function just gives us one of the standard basis
vectors in p (Rn )? . This is a constant function, so its certainly smooth, and
so this is a p-form.
A general p-form on U is given by some smooth function
: U
p
n ?
(R ) , and we can write it as
X
i d
xi1 ... d
xi p
i
where i runs over all correctly-ordered p-tuples i = {i1 < ... < ip } and each
coefficient
i is a smooth function from U to R. If we wanted to we could
104
=
1 d
x1 + ... +
n d
xn
and
= 1 d
x1 + ... + n d
xn
=
(
i j
j i )d
xi d
xj
i<j
We need to check that F ? is smooth, but we should first note that it follows
immediately from Proposition 8.9(iii) that
F ? ( ) = F ? () F ? ()
(8.11)
for any two differential forms and on Y . Now it is easy to check that F ?
is smooth, because in co-ordinates we can write as a linear combination of
105
and
F ? dy = sin dr + r cos d
2 : U2 p (Rn )?
1 : U1 p (Rn )?
2 = ?12
1
on the open set f2 (U1 U2 ) U2 . As for vector fields and one-forms, we
could take this transformation law as the definition of a p-form if we wished.
We could write this transformation law explicitly in terms of the Jacobian
matrix D12 |f2 (x) , but it gets rather complicated. However, in the special case
106
of n-forms, its very easy. If is an n-form, then in each chart just becomes
smooth functions:
1 : U1 R
and
2 : U2 R
By our observation (8.5), on the overlap these functions are related by:
h
h
d
x1 + ... +
d
xn
x1
xn
So we have an operator:
d : 0 (X) 1 (X)
h 7 dh
What we want to do is extend this to an operator
d : p (X) p+1 (X)
for all p.
Well begin by assuming X is an open set U Rn , and well let x1 , ..., xn
be the co-ordinate functions on U . Suppose we have a p-form
which has
only one non-zero component, so
=
i d
xi1 ... d
xi p
where i = (i1 < ... < ip ) is a single correctly-ordered p-tuple, and
i
C (U ). Then we define:
d
= d
i d
xi1 ... d
xi p
n
X
i
d
xj d
xi1 ... d
xi p
=
xj
j=1
107
(8.14)
Some of the terms in this sum will be zero, since when j is equal to one of
the it then that wedge product of one-forms is zero. So there will be one
(potentially) non-zero component of d
for each j which does not appear in
the p-tuple i, and if we want to write it in terms of our standard basis then we
pick up a 1 when we apply the permutation which returns the (p + 1)-tuple
(j, i1 , ..., ip ) to its correct order.
The expression (8.14) is linear in
i , so we can extend it to a linear
operator:
d : p (U ) p+1 (U )
Together, these operators are called the exterior derivative, or the de Rham
differential.
Example 8.15. Let X = R3 , with co-ordinates x, y and z. If we have a
one-form
=
2 dy for some
2 C (R2 ), then:
d
=
2
dx dy
dy dz
x
z
More generally, if
=
1 dx +
2 dy +
3 dz
then:
d
=
2
1
3
1
3
2
dxdy +
dxdz +
dydz
x
y
x
z
y
z
You might recognise this formula - if we flip the sign of middle term, which
we can do by deciding to write things in terms of dz dx instead of dx dz,
then this is the formula for the curl operator which turns a vector field
on R3 into another vector field on R3 . However its more natural to interpret
it as an operator that turns one-forms into 2-forms.
The exterior derivative has many nice properties.
Proposition 8.16.
(8.17)
A good way to remember the sign in part (ii) is to pretend that the symbol
d behaves a bit like a one-form, so if we want to permute it past the p-form
then we pick up a sign (1)p . Notice that (8.17) is formally similar to the
product rule for derivations, and in fact d is indeed a derivation in a more
general sense.
Proof. (i) By linearity its enough to prove the result for an
which has
a single component
i for some i. Applying the formula (8.14) twice,
we get the (p + 2)-form:
n X
n
X
2
i
d
xm d
xj d
xi1 ... d
xi p
d(d
) =
m
j
m=1 j=1
This sum has a (potentially) non-zero term for every pair m, j such that
m 6= j and neither m nor j appear in i. However, the double partial
2
derivative xmixj is symmetric in m and j, and the wedge product
d
xm d
xj is antisymmetric in m and j, so these terms cancel in pairs.
(ii) Firstly suppose that
and are just zero-forms, i.e. elements of
C (U ). Then (8.17) says that
= d
+
d(
)
d
(since for zero-forms the wedge-product is ordinary point-wise multiplication), and this is true by the product rule for partial differentiation.
Now let
C (U ) be a zero-form, and let be the constant q-form
d
xi1 ... d
xiq . Then the formula (8.14) says that d = 0, and it also
says that
d(
d
xi1 ... d
xiq ) = d
d
xi1 ... d
xi q
so this special case of (8.17) is also true.
Now let
be a p-form with a single component, and be a q-form with
a single component, so:
and
= j d
xj ... d
xj q
=
i d
xi ... d
xi p
1
with
i , j C (U ). Then:
=d
d(
)
i j d
xi1 ... d
xip d
xj1 ... d
xj p
= j d
i +
i dj d
xi1 ... d
xip d
xj1 ... d
xj p
= (d
i d
xi1 ... d
xip ) (j d
xj1 ... d
xjq )
+ (1)p (
i d
xi1 ... d
xip ) (dj d
xj1 ... d
xj q )
= d
+ (1)p
d
109
7.12, so d(F d
xk ) = 0 by part (i) of this proposition. Then repeatedly
applying part (ii) of this proposition shows that:
d(F ?
) = d(
i F ) (F ? d
xi1 ) ... (F ? d
xi p )
Hence
d(F ?
) = (F ? d
i ) (F ? d
xi1 ) ... (F ? d
xi p )
?
= F d
i d
xi1 ... d
xi p
= F ? d
as required.
Now we want to define the exterior derivative on an arbitrary manifold X.
Obviously, when we work in co-ordinates it should reduce to the operations
that weve just defined. In fact, this requirement is a valid way to define d
on X.
Fix a p-form p (X) . In any chart (U, f ), it turns into a p-form
: U p (Rn )?
on U . So in any chart, we can form a (p + 1)-form:
d
: U p+1 (Rn )?
So we have a rule which produces, for any chart (U, f ), a (p + 1)-form on U .
We claim that this object is actually a (p + 1)-form on X, in the physicists
sense, i.e. we claim that it obeys the correct transformation law when we
change co-ordinates. This means that there really is a (p + 1)-form
d p+1 (X)
such that when we write d in a chart (U, f ) we get the corresponding d
.
110
2 = ?12
where the two charts overlap. But by Proposition 8.16(iii), we have
1
d
2 = ?12 d
and this is the correct transformation law for a (p + 1)-form on X. This
proves our claim.
This means that on any manifold X, and any p, we have a exterior derivative:
d : p (X) p+1 (X)
Futhermore its immediate that the three properties listed in Proposition
8.16 all hold, since each of them can be checked in co-ordinates.
The main use for the exterior derivative is to define the de Rham cohomology of a manifold, which unfortunately is not a part of this course. However,
we will see one application of d in the next section.
Integration
C (U ), we can
If we have an open set U Rn , and a smooth function h
try to compute the multiple integral:
Z
d
h
x1 ...d
xn
x1 ...d
xn = (h F ) det DF d
y1 ...d
yn
(9.1)
111
The factor | det DF |, the absolute value of the determinant of the Jacobian
matrix of F , keeps track of how volume gets distorted when we change variables.
This is strikingly similar to the the way
R that n-forms behave. If we decide
that the symbols after the integral sign U are really the n-form
d
=h
x1 ... d
xn n (U )
then we have
F ) det(DF ) d
F ?
= (h
y1 ... d
yn
R
which is almost
the same
as the symbols after V . The only difference is the
occurence of det(DF ) instead of det(DF ), and well deal with this shortly.
This is very strong evidence that the correct thing to integrate over U is not
functions but n-forms, and we should define:
Z
Z
d
=
h
x1 ...d
xn
9.1
Orientations
In this section were going to solve problem (b), the problem of whether the
integral of an n-form over a co-ordinate chart is co-ordinate independent.
Let X be a manifold, and let n (X) be an n-form on X. Suppose
we pick two charts U1 and U2 , and let U = U1 U2 . On U we have two
co-ordinate systems, with co-domains f1 (U ) U1 and f2 (U ) U2 , and
=h
x1 ... d
xn
112
n (f1 (U ))
Then we know from the transition law for n-forms that when we write in
the second co-ordinates we will get:
12 ) det(D12 ) d
?12
= (h
y1 ... d
yn n (f2 (U ))
R
We would like to be able to define the integral
R U , by evaluating it in corodinates. In the first co-ordinates this gives f1 (U )
, and lets assume that
this integral converges. Then by the change-of-variables formula (9.1) we
have that:
Z
Z
12 ) |det(D12 )| d
=
(h
y1 ...d
yn
f1 (U )
f2 (U )
Z
?12
(9.2)
f2 (U )
If we knew that det D12 was always positive then this would be an equality.
In fact, since 12 is a diffeomorphism, we know that det D12 is never zero.
So if we assume that U is connected then det D12 must be either always
positive or always negative, and we have an equality up to an overall factor
of 1.
R
If we can solve this sign issue, then we can evaluate U in either coordinates and get the same answer. The solution to the signs is provided by
the idea of an orientation.
Definition 9.3. A volume form on a manifold X is an n-form n (X)
such that is not zero at any point. If there exists a volume form on X then
we say that X is orientable.
Example 9.4. If U Rn is an open set, then
0 = d
x1 ... d
xn n (U )
is a volume form, so Rn is orientable. Well call
0 the standard volume form
0 for some h
C (U )
on U . Every other volume form on U is of the form h
which is never zero.
Since the vector bundle n T ? X has rank 1, asking for X to be orientable
is equivalent to asking for n T ? X to be a trivial vector bundle, by Proposition 6.19. Also, if weve fixed a volume form on X, then any n-form
n (X) can be written as = h for some h C (T 1 ). If h is never
zero then will be another volume form (and vice-versa).
113
Example 9.5. Let X = T 1 , and lets use our usual atlas with two charts
where:
U1 = (0, 1)
and
U2 = ( 21 , 21 )
Since D21 = 1 at all points, a one-form 1 (T 1 ) must be expressed in
this atlas as
1 d
1 = h
x 1 (U1 )
and
2 d
2 = h
y 1 (U2 )
where:
2 (x) = h
1 (x)
h
x (0, 21 ), and
2 (x) = h
1 (x + 1) x ( 1 , 0)
h
2
1 1 and h
2 1, and this defines a volume form
In particular we may set h
1
1
1
(T ). Hence T is orientable.
1 and h
2 are
Notice that, for any one-form 1 (T 1 ), the data of h
1
exactly the data of a function h C (T ), and we have that = h.
We saw in Example 6.21 that S 1 (and hence T 1 ) is parallelizable, and
its not hard to show in general that if X is parallelizable then X is also
orientable. However, being orientable is a much weaker condition, and in
practice most manifolds that we care about are orientable.
Proposition 9.6. Let X be an orientable manfold, and let
Z = h1 (y) X
be a level set of some function h C (X) at a regular value y R. Then
Z is orientable.
In particular if we set X = Rn+1 and Z = S n then we see that S n is
orientable, for any n.
Proof. Let dim Z = n, and fix a point z Z, so the volume form at this
point gives us a non-zero element |z n Tz? X. We saw in Lemma 5.15
that the tangent space Tz Z is the subspace of Tz X given by the kernel of the
linear map Dh|z : Tz Z Tz R
= R. Let n Tz X be any vector such that
Dh|z (n) = 1. Then we can define an element 0 |z n1 Tz? Z by declaring
that
0 |z : (v1 , ..., vn1 ) 7 |z (v1 , ..., vn1 , n) R
for any vectors v1 , ..., vn1 Tz Z. This map 0 |z is automatically (n 1)linear and antisymmetric, so it is indeed an element of n1 Tz? Z. Futhermore
114
we claim that its independent of our choice of n. To see this, recall that
n1 Tz? Z is only 1-dimensional, so if we pick a basis e1 , ..., en1 for Tz Z then
0 |z is determined by the single real number :
0 |z (e1 , ..., en1 ) = |z (e1 , ..., en1 , n)
(9.7)
This number will not change if we change n by adding on any linear combination of the ei s, because |z is anti-symmetric and linear in each argument.
However, any vector in Dh|1
z (1) must differ from n by some linear combi0
nation of the ei s, so |z is indeed independent of our choice of n.
Also, the number (9.7) cannot be zero, because the vectors e1 , ..., en1 , n
form a basis of Tz X, and we know that |z is not zero. Therefore 0 |z is not
the zero element of n1 Tz? Z.
So for every point z Z, we have constructed a non-zero element 0 |z
n1 Tz? Z. If we can show that these elements vary smoothly, then we have
found a volume form 0 on Z. So we need to look at this construction in
co-ordinates.
We can assume that Z is the level set of y = 0, since we can always
replace h by h y. Then for any point z Z, we know that we can find a
chart (U, f ) containing z such that when we write h in this chart it is just the
= xn on U . In such a chart, the submanifold Z becomes
last co-ordinate h
n1
f (Z U ) = R
U , and we may choose our vector n to be the tangent
= g d
x1 ... d
xn
for some g C (U ). It follows that 0 , in these co-ordinates, is given by
0 = g|xn =0 d
x1 ... d
xn1
which is indeed a smooth (n 1)-form on f (Z U ).
With a little more work, this proposition can be generalised to level sets
(at regular values) of smooth functions h : X Y , where Y is any other
orientable manifold. However, it is not true that any submanifold of an
orientable manifold is orientable.
Now let X be an orientable manifold, and lets fix a volume form on
X. Pick a chart (U, f ) on X. In this chart becomes a volume form
n (U )
115
1 = h
0 n (U1 )
and
2 = h
0 n (U2 )
1 and h
2 are positive everywhere since both charts are oriented.
where both h
For a point x U1 U2 we have
by the transformation law for n-forms (8.13). Hence det D12 |f2 (x) > 0.
This solves part of our problem of defining integration on manifolds. Lets
return to the situation we discussed at the beginning of this section, where
we have an n-form n (X), and a region U X which is the intersection
of two charts, so it has two co-ordinate systems f1 and f2 . If we assume that
X is oriented, and that both charts are oriented, then the transition function
12 satisfies det D12 > 0 at all points. Consequently the formula (9.2) is an
equality,
so (assuming that the integral converges) we can define the integral
R
using either co-ordinate system and we will get the same answer.
U
The other part of the problem (part (a)) is about the convergence. Well
solve this in the next section.
9.2
Defining integration
as some real number. However this is not going to work in general, because
integrals do not always converge. For example if we take X = R, and
1 (R) to be the constant one-form dx, then we are trying to evaluate
Z
1 dx
since this integral converges. Now lets prove that this definition is independent of our choice of chart.
Proposition 9.11. Let (U1 , f1 ) be an oriented chart such that vanishes
outside of some compact subset W1 U1 . Let (U2 , f2 ) be another oriented
chart such that vanishes outside of some compact subset W2 U2 . Let
1 n (U1 ) and
2 n (U2 ) be the n-forms that we get by writing in the
two charts. Then we have:
Z
Z
1 =
2
1
U
2
U
R
X
R.
Proof. Let U = U1 U2 , and W = W1 W2 . Then W is a compact subset of
U , and vanishes outside of W . Consequently
Z
Z
Z
Z
1
and
2 =
1 =
1
U
2
U
f1 (U )
f2 (U )
since
1 vanishes outside f1 (U ) and
2 vanishes outside f2 (U ), and all these
integrals converge. Weve now reduced to the situation considered in the
previous section, and the formula (9.2) shows that
Z
Z
1 =
2
f1 (U )
f2 (U )
since
2 = 12
1 and both charts are oriented.
We now want to consider integrating arbitrary n-forms. This means that
we have to put some restriction on X, and the correct restriction is to insist
that X itself is compact.
On a compact manifold, there is a way to chop-up an arbitrary n-form
into a finite number of bump forms. Then we can then integrate each piece,
and add the answers together. The chopping-up step is done with the
following gadget:
Definition 9.12. Let X be a manifold. A partition-of-unity on X is a
finite set of functions 1 , ..., r C (X) with the following two properties:
118
(i) For each i the function i 0 (X) is a bump form, so there exists
a chart (Ui , fi ) and a compact subset Wi Ui such that i vanishes
outside of Wi .
(ii)
1 + ... + r 1
Again we are not quite using standard terminology here - everyone would
agree that a partition-of-unity must satisify property (ii), but some people
might vary property (i), and/or allow the set of functions to be infinite.
Proposition 9.13. If X is compact then a partition-of-unity exists on X.
In fact its fairly easy to show that a (finite) partition-of-unity can only
exist if X is compact, so the proposition is really if-and-only-if.
Proof. For any point x X, we can find a bump function x which is
constantly equal to 1 on some open neighbourhood Vx of x, never negative, and vanishes outside of some compact set which is contained within
a chart. Choose this data of x and Vx and for each point x. The open
sets {Vx , x X} form an open cover of X, so since X is compact there is
some finite subcover {Vx1 , ..., Vxr }. Let xr , ..., xr be the corresponding set
of bump functions. Then the sum x1 + ... + xr is strictly positive at all
points of X, since no term is negative and at all points at least one term is
equal to 1. Hence we can define
i =
x1
xi
C (X)
+ ... + xr
We just need to check that this definition is independent of which partitionof-unity we chose.
119
X,
X,
b
bj i
X,
X,
b
i,j
If we swap and
b then we get the same answer, which proves the proposition.
So we just need to prove the claim. Suppose is a bump-form, and (U, f )
is an oriented chart such that vanishes outside of a compact subset W U .
Now let be any a partition-of-unity.R Each bump-form i also vanishes
outside of W , so we may evaluate each X i using the chart (U, f ). Writing
everything in these co-ordinates, we have that
X
d
x1 ... d
xn =
i d
x1 ... d
xn
i
and
over U is a linear operation, we see that
R
P R since integrating
i X i = X, .
R
X
First we need a partition-of-unity. Let
b1 C (0, 1) be some function
which is never negative, constantly equal to 1 in some interval containing
120
1
,
2
T1
T1
b1 (x)b
h(x) dx
1 h =
T1
Z
2 h =
T1
1
2
b2 (x)b
h(x) dx =
b2 (x)b
h(x) dx
since b
h and
b2 are periodic. So adding the two terms together gives:
Z 1
Z 1
Z
b
b
h(x) dx
h =
b1 (x) +
b2 (x) h(x) dx =
T1
Theres an easy general observation we can make here: if X is any compact manifold and n (X) is a volume form, then if we use the orientation
[] on X we must have
Z
>0
X
since the integral will be a sum of strictly positive terms. This means that
integration over X defines a surjective linear map
Z
: n (X) R
X
R
X
to take any
9.3
Stokes Theorem
h
dx
x
R
also vanishesR outside Rof this interval, so the integral R dh certainly converges,
r
and in fact R dh = r dh. So by the fundamental theorem of calculus, we
have
Z r
Z
h
dx = h(r) h(r) = 0
dh =
r x
R
since h(r) = h(r) = 0.
Lets generalise this observation to higher dimensions. If we work on Rn ,
then the objects that we can integrate are n-forms, so we must replace h by
an (n 1)-form. For example, lets work on R3 , and consider a 2-form
= 1 dy dz
which only has one non-zero component. Then:
d =
1
dx dy dz
x
Suppose that vanishes outside some cube, i.e. the function 1 C (R3 )
is zero unless x, y, z [r, r]. Then we have that
Z r Z r Z r
Z
1
dx dy dz
d =
r
r
r x
R3
Z r Z r
=
1 (r, y, z) 1 (r, y, z) dy dz
r
=0
since 1 (r, y, z) = 1 (r, y, z) = 0 for any values of y and z.
R
If has more than one component then its still true that R R3 d = 0,
since we can evaluate each component individually (both d and are linear)
and each piece will be zero. We can also perform this argument in exactly
the same way in higher dimensions - the only problem is keeping track of the
notation! Formally:
Lemma 9.16. Let U Rn be an open set, and let n1 (U ) be an
(n 1)-form that vanishes outside some compact subset W U . Then:
Z
d = 0
122
[r,r]n
=
T1
b
h(x) dx
0
where b
h C (R) is the periodic function corresponding to h CR (T 1 ). If
= dg for some g C (T 1 ), then Stokes Theorem says that T 1 = 0.
R1
Therefore, if 0 h(x)
dx 6= 0, then there cannot exist a g C (T 1 ) such that
= dg. In particular there is no g such that = dg.
R1
The converse to this statement is also true. If 0 b
h(x) dx = 0, then the
function
gb : R R
Z x
b
x 7
h(y) dy
0
123
124
Topological spaces
If X is a set we let PX denote the power set of X, i.e. the set of all subsets
of X.
Definition A.1. Let X be a set. A topology on X is a collection
T PX
of subsets of X, satisfying the list of axioms below. We refer to elements of
T as open sets. The axioms are:
(i) The empty subset is open, and the whole of X is open.
(ii) The intersection of two open sets is open.
(iii) Given any collection of open sets, their union is also open.
Axiom (ii) implies that the intersection of any finite collection of open sets
is open. Axiom (iii) applies to any collection of open sets, including infinite
ones. If we have chosen a topology on X, then we call X a topological
space.
A subset V of a topological space X is called closed iff its complement
c
V = X \V is open. Note that most subsets of X are neither open nor closed.
Example A.2. Let X = Rn , equipped with the usual (Euclidean) norm.
For a point x Rn , and a real number r R0 , the open ball around x of
radius r is the set:
B(x, r) = {y Rn ; |y x| < r}
We declare that a subset U Rn is open iff for any point x U there exists
some > 0 such that:
B(x, ) U
Equivalently, we can say that a subset U X is open iff U can be written as
a union of some collection of open balls. It is easy to prove that this defines
a topology on Rn .
Definition A.3. Let X and Y be topological spaces, and let f be a function:
f :XY
We say that f is continuous iff whenever U Y is an open set then its
pre-image
f 1 (U ) X
is also open. Equivalently, we can require that the pre-image of every closed
set is closed.
125
126
128
129