6.1 Review: Differential Forms On R
6.1 Review: Differential Forms On R
6.1 Review: Differential Forms On R
Differential forms
where wi1 ...ik 2 C• (U) are functions, and the indices are numbers
1 i1 < · · · < ik m.
Let W k (U) be the vector space consisting of such expressions, with pointwise addi-
tion. It is convenient to introduce a short hand notation I = {i1 , . . . , ik } for the index
set, and write w = ÂI wI dxI with
The dxI are just formal expressions; at this stage they do not have any particular
meaning. They are used, however, to define an associative product operation
dxi ^ dx j = dx j ^ dxi
for all i, j; in particular dxi ^ dxi = 0. In turn, using the product structure we may
define the exterior differential
102 6 Differential forms
⇣ ⌘ m
∂ wI
d : W k (U) ! W k+1 (U), d  wI dxI =   i dxi ^ dxI . (6.1)
I i=1 I ∂ x
d d = 0,
which vanishes by equality of mixed partials ∂ ∂xiw∂Ix j = ∂ ∂x jw∂Ixi . (We have dxi ^ dx j =
dx j ^ dxi , but the coefficients in front of dxi ^ dx j and dx j ^ dxi are the same.) t u
Exercise 75.
(a) A 0-form on R3 is simply a smooth function f 2 W 0 (R3 ). Use the definition
of the exterior differential above to compute the resulting 1-form d f .
(b) A general 1-form w 2 W 1 (R3 ) is an expression
w = f dx + gdy + hdz
w = a dy ^ dz + b dz ^ dx + c dx ^ dy,
curl(grad( f )) = 0, div(curl(F)) = 0
The support supp(w) ✓ U of a differential form is the smallest closed subset such
that w vanishes on U\supp(w). Suppose w 2 W m (U) is a compactly supported form
of top degree k = m. Such a differential form is an expression
6.2 Dual spaces 103
w = f dx1 ^ · · · ^ dxm
Note that we can regard w as a form on all of Rm , due to the compact support condi-
tion.
Our aim is now to define differential forms on manifolds, beginning with 1-
forms. Even though 1-forms on U ✓ Rm are identified with functions U ! Rm ,
they should not be regarded as vector fields, since their transformation properties
under coordinate changes are different. In fact, while vector fields are sections of the
tangent bundle, the 1-forms are sections of its dual, the cotangent bundle. We will
therefore begin with a review of dual spaces in general.
ha, vi := a(v).
This pairing notation emphasizes the duality between a and v. In the notation a(v)
we think of a as a function acting on elements of E, and in particular on v. However,
one may just as well think of v as acting on elements of E ⇤ by evaluation: v(a) =
a(v) for all a 2 E ⇤ . This symmetry manifests notationally in the pairing notation.
Let e1 , . . . , er be a basis of E. Any element of E ⇤ is determined by its values on
these basis vectors. For i = 1, . . . , r, let ei 2 E ⇤ (with upper indices) be the linear
functional such that ⇢
i i 0 if i 6= j,
he , e j i = d j =
1 if i = j.
The elements e1 , . . . , er are a basis of E ⇤ ; this is called the dual basis. The element
a 2 E ⇤ is described in terms of the dual basis as
r
a= Â a j e j, a j = ha, e j i.
j=1
⇤ For possibly infinite-dimensional vector spaces, the dual space E ⇤ is not isomorphic to E,
in general.
† In physics, one also uses the Dirac bra-ket notation ha| vi := a(v); here a = ha| is the
‘bra’ and v = |vi is the ‘ket’.
104 6 Differential forms
Notice the placement of indices: In a given summation over i, j, . . ., upper indices are
always paired with lower indices.
Remark 6.1. As a special case, for Rr with its standard basis, we have a canonical
identification (Rr )⇤ = Rr . For more general E with dim E < •, there is no canonical
isomorphism between E and E ⇤ unless more structure is given.
Exercise 76. Let V be a finite dimensional real vector space equipped with an
inner product h·, ·i. Every vector v 2 V determines a linear transformation Av by
v 7! hv, ·i.
Given a linear map R : E ! F between vector spaces, one defines the dual map
R⇤ : F ⇤ ! E ⇤
(note the direction), by setting
hR⇤ b , vi = hb , R(v)i
for b 2 F ⇤ and v 2 E. This satisfies (R⇤ )⇤ = R, and under the composition of linear
maps,
(R1 R2 )⇤ = R⇤2 R⇤1 .
In terms of basis e1 , . . . , er of E and f1 , . . . , fs of F, and the corresponding dual bases
(with upper indices), a linear map R : E ! F is given by the matrix with entries
Ri j = h f j , R(ei )i,
while R⇤ is described by the transpose of this matrix (the roles of i and j are re-
versed). Namely,‡
‡ In bra-ket notation, we have Ri j = h f j |R |ei i, and
|Rei i = R|ei i = Â | f j ih f j |R |ei i, hR⇤ ( f j )| = h( f j )|R = h f j |R |ei ihei |
j
6.3 Cotangent spaces 105
s r
R(ei ) = Â Ri j f j , R⇤ ( f j ) = Â Ri j f i .
j=1 i=1
Thus,
(R⇤ ) j i = Ri j .
Definition 6.1. The dual of the tangent space Tp M of a manifold M is called the
cotangent space at p, denoted
Elements of Tp⇤ M are called cotangent vectors, or simply covectors. Given a smooth
map F 2 C• (M, N), and any p 2 M we have the cotangent map
(d f ) p 2 Tp⇤ M, h(d f ) p , vi = v( f ).
§ Note that this is actually the same as the tangent map Tp f : Tp M ! T f (p) R = R.
106 6 Differential forms
The basis of the dual space Tp⇤U, dual to the basis (6.3), is given by the differentials
of the coordinate functions:
Indeed,
D ∂ E ∂
(dxi ) p , = (xi ) = d i j
∂xj p ∂xj p
as required. For f 2 C• (M), the coefficients of (d f ) p = Âi h(d f ) p , ei iei are deter-
mined as D ∂ E ∂ ∂f
(d f ) p , = (f) = .
∂xj p ∂xj p ∂xj p
Thus,
m
∂f
(d f ) p = Â i (dxi ) p .
i=1 ∂ x p
∂F j
(D p F)i j = (p)
∂ xi
for i = 1, . . . , m, j = 1, . . . , n. We have:
n
∂ ∂
(Tp F)(
∂ xi p
)= Â (D p F)i j ∂yj F(p)
,
j=1
hence dually
m
(Tp F)⇤ (dy j )F(p) = Â (D p F)i j (dxi ) p . (6.4)
i=1
Thought of as matrices, the coefficients of the cotangent map are the transpose of the
coefficients of the tangent map.
6.4 1-forms
Similar to the definition of vector fields, one can define co-vector fields, more com-
monly known as 1-forms: Collections of covectors a p 2 Tp⇤ M depending smoothly
6.4 1-forms 107
on the base point. One approach of making precise the smooth dependence on the
base point is to endow the cotangent bundle
G
T ⇤M = Tp⇤ M.
p
(disjoint union of all cotangent spaces) with a smooth structure, and require that the
map p 7! a p be smooth. The construction of charts on T ⇤ M is similar to that for
the tangent bundle: Charts (U, j) of M give cotangent charts (T ⇤U, T ⇤ j 1 ) of T ⇤ M,
using the fact that T ⇤ (j(U)) = j(U)⇥Rm canonically (since j(U) is an open subset
of Rm ). Here T ⇤ j 1 : T ⇤U ! T ⇤ j(U) is the union of inverses of all cotangent maps
Tp⇤ j : Tj(p)
⇤ j(U) ! T ⇤U.
p
Exercise 79. Carry out this construction to prove that T ⇤ M is naturally a 2m-
dimensional manifold.
A second approach is observe that in local coordinates, 1-forms are given by
expressions Âi fi dxi , and smoothness should mean that the coefficient functions are
smooth.
We will use the following (equivalent) approach.
Definition 6.3. A 1-form on M is a linear map
a : X(M) ! C• (M), X 7! a(X) = ha, Xi,
which is C• (M)-linear in the sense that
a( f X) = f a(X)
for all f 2 C• (M), X 2 X(M). The space of 1-forms is denoted W 1 (M).
Let us verify that a 1-form can be regarded as a collection of covectors:
Lemma 6.2. Let a 2 W 1 (M) be a 1-form, and p 2 M. Then there is a unique covector
a p 2 Tp⇤ M such that
a(X) p = a p (X p )
for all X 2 X(M).
(We indicate the value of the function a(X) at p by a subscript, just like we did for
vector fields.)
Proof. We have to show that a(X) p depends only on the value of X at p. By consid-
ering the difference of vector fields having the same value at p, it is enough to show
that if X p = 0, then a(X) p = 0. But any vector field vanishing at p can be written as a
finite sum X = Âi fiYi where fi 2 C• (M) vanish at p. ¶ By C• -linearity, this implies
that
a(X) = a(Â fiYi ) = Â fi a(Yi )
i i
vanishes at p. t
u
¶ ∂
For example, using local coordinates, we can take the Yi to correspond to ∂ xi
near p, and
the fi to the coefficient functions.
108 6 Differential forms
d f 2 W 1 (M),
a = Â fi dgi
i
where fi , gi 2 C• (M).
Let us examine what the 1-forms are for open subsets U ✓ Rm . Given a 2 W 1 (U),
we have
m
a = Â ai dxi
i=1
⌦ ↵
with coefficient functions ai = a, ∂∂xi 2 C• (U). (Indeed, the right hand side takes
on the correct values at any p 2 U, and is uniquely determined by those values.)
General vector fields on U may be written
m
∂
X= Â X j ∂xj
j=1
(to match the notation for 1-forms, we write the coefficients as X i rather than ai , as
we did in the past), where the coefficient functions are recovered as X j = hdx j , Xi.
The pairing of the 1-form a with the vector field X is then
m
ha, Xi = Â ai X i .
i=1
F ⇤ : C• (M 0 ) ! C• (M), f 7! F ⇤ ( f ) := f F.
Let F 2 C• (M, N) be a smooth map. Recall that for vector fields, there is no
general ‘push-forward’ or ‘pull-back’ operations, unless F is a diffeomorphism. For
1-forms the situation is better. Indeed, for any p 2 M one has the dual to the tangent
map
Tp⇤ F = (Tp F)⇤ : TF(p)
⇤
N ! Tp⇤ M.
For a 1-form b 2 W 1 (N), we can therefore define
(F ⇤ b ) p := (Tp⇤ F)(bF(p) ).
(a) Compute F ⇤ (du) and F ⇤ (v cos udv). Compute F ⇤ (v cos udu + sin udv) by
using Lemma 6.1.
(b) Verify Lemma 6.1 by computing dg, F ⇤ (dg), as well as F ⇤ g and d(F ⇤ g).
Recall once again that while F 2 C• (M, N) induces a tangent map T F 2 C• (T M, T N),
there is no natural push-forward operation for vector fields. By contrast, for cotan-
gent bundles there is no naturally induced map from T ⇤ N to T ⇤ M (or the other way),
yet there is a natural pull-back operation for 1-forms!
In the case of vector fields, rather than working with ‘F⇤ (X)’ one has the notion
of related vector fields, X ⇠F Y . We know that 1-forms act on vector fields, how do
they act on related vector fields?
Exercise 83. Show that for any related vector fields X ⇠F Y , and b 2 W 1 (N),
(F ⇤ b )(X) = F ⇤ (b (Y )).
(Notice once again the different notions of pull-back that we are using.)
Exercise 84. Recall that for a given vector field on a manifold X 2 X(M), the
smooth curve g 2 C• (J, M) (where J ✓ R) is a solution curve iff
∂
gX.
∂t
Let g : R ! R2 be given by
6.6 Integration of 1-forms 111
t 7! (cost, sint).
g ⇤ a = f (t)dt
Definition 6.5. Given a smooth path g : [a, b] ! M, we define the integral of a 1-form
a 2 W 1 (M) along g as
Z Z b
a= g ⇤ a.
g a
The fundamental theorem of calculus has the following consequence for mani-
folds. It is a special case of Stokes’ theorem.
Proposition 6.2. Let g : [a, b] ! M be a smooth path, with g(a) = p, g(b) = q. For
any f 2 C• (M), we have Z
d f = f (q) f (p).
g
In particular, the integral of d f depends only on the end points of the path, rather
than the path itself.
Proof. We have
∂ ( f g)
g ⇤ d f = dg ⇤ f = d( f g) = dt.
∂t
Integrating from a to b, we obtain, by the fundamental theorem of calculus, f (g(b))
f (g(a)). tu
112 6 Differential forms
Exercise 85. Given a diffeomorphism k : [c, d] ! [a, b] one defines the corre-
sponding reparametrization
g k : [c, d] ! M.
a = y2 ex dx + 2yex dy 2 W (R2 ).
g : [0, 1] ! M, t 7! (sin(pt/2),t 3 ).
∂ ai ∂aj
= ,
∂xj ∂ xi
In multivariable calculus one learns that this condition is also sufficient, provided U
is simply connected (e.g., convex). Using the exterior differential of forms in W 1 (U),
this condition becomes da = 0. Since a is a 1-form, da is a 2-form. Thus, to obtain
a coordinate-free version of the condition, we need higher order forms.
6.7 k-forms
To get a feeling for higher degree forms, and constructions with higher forms, we
first discuss 2-forms.
6.7 k-forms 113
6.7.1 2-forms.
Here skew-symmetry means that a(X,Y ) = a(Y, X) for all vector fields X,Y , while
C• (M)-bilinearity means
1 m
2 i,Â
w= wi j dxi ^ dx j = Â wi j dxi ^ dx j .
j=1 i< j
6.7.2 k-forms
Here, skew-symmetry means that a(X1 , . . . , Xk ) changes sign under exchange of any
two of its elements. For example, a(X1 , X2 , X3 , . . .) = a(X2 , X1 , X3 , . . .). More gen-
erally, if Sk is the group of permutations of {1, . . . , k}, and sgn(s) is the sign of a
permutation s 2 Sk (+1 for an even permutation, 1 for an odd permutation) then
a p : Tp M ⇥ · · · ⇥ Tp M ! R,
for all p 2 M.
If a1 , . . . , ak are 1-forms, then one obtains a k-form a =: a1 ^ . . . ^ ak by ‘wedge
product’.
Exercise 89. Show that the wedge-product above indeed defines a k-form.
Exercise 91.
(a) Show that Definition 6.8 is consistent with our previous definition of the
wedge-product of two 1-forms (Equation (6.5)).
(b) Show that Definition 6.8 indeed defines a (k + l)-form.
Exercise 92.
(a) For a, b , r 2 W 2 (M), and T,U,V,W, X,Y, Z 2 X(M), compute
(a ^ b )(W, X,Y, Z)
and
(a ^ b ) ^ r(T,U,V,W, X,Y, Z).
(b) For a, b 2 W 3 (M), and T,U,V,W, X,Y, Z 2 X(M), compute
Exercise 93.
(a) Prove that the wedge product is graded commutative: If a 2 W k (M) and
b 2 W l (M) then
a ^ b = ( 1)kl b ^ a.
116 6 Differential forms
(b) Prove that the wedge product is associative: Given ai 2 Wki (M) we have
(a1 ^ a2 ) ^ a3 = a1 ^ (a2 ^ a3 ).
So, we may in fact drop the parentheses when writing wedge products.
Recall that we have defined the exterior differential on functions by the formula
(d f )(X) = X( f ). (6.6)
Theorem 6.1. There is a unique collection of linear maps d : W k (M) ! W k+1 (M),
extending the map (6.6) for k = 0, such that d(d f ) = 0 and satisfying the graded
product rule,
d(a ^ b ) = da ^ b + ( 1)k a ^ db (6.7)
for a 2 W k (M) and b 2 W l (M). This exterior differential satisfies d d = 0.
Proof. Suppose first that such an exterior differential is given. Then d is local, in the
sense that for any open subset U ✓ M the restriction (da)|U depends only on a|U ,
or equivalently (da)|U = 0 when a|U = 0. Indeed, if this is the case and p 2 U, we
may choose f 2 C• (M) = W 0 (M) such that f vanishes on M\U and f | p = 1. Then
f a = 0, hence the product rule (6.7) gives
0 = d( f a) = d f ^ a + f da.
Conversely, we may use this explicit formula (cf. (6.1)) to define da|U for a coor-
dinate chart domain U; by uniqueness the local definitions on overlas of coordinate
chart domains agree. Proposition 6.1 shows that (dda)|U = 0, hence it also holds
globally. tu
6.8 Lie derivatives and contractions 117
Exercise 94. Find the exterior differential of each of the following forms on
R3 (with coordinates (x, y, z)).
(a) a = y2 ex dy + 2yex dx.
(b) b = y2 ex dx + 2yex dy.
2
(c) r = ex y sin zdx ^ dy + 2 cos(z3 y)dx.
sin exy cos sin z3 x
(d) w = 1+(x+y+z) 4 +(7xy)6 dx ^ dy ^ dz.
{a 2 W k (M)| a is closed }
H k (M) = .
{a 2 W k (M)| a is exact }
It turns out that whenever M is compact (and often also if M is non-compact), H k (M)
is a finite-dimensional vector space. The dimension of this vector space
is called the k-th Betti number of M; these numbers are important invariants of M
which one can use to distinguish non-diffeomorphic manifolds. For example, if M =
CPn one can show that
bk (CPn ) = 1 for k = 0, 2, . . . , 2n
bk (Sn ) = 1 for k = 0, n
while bk (Sn ) = 0 for all other k. Hence CPn cannot be diffeomorphic to S2n unless
n = 1.
iX a 2 W k 1
(M)
by contraction: Thinking of a as a multi-linear form, one simply puts X into the first
slot:
(iX a)(X1 , . . . , Xk 1 ) = a(X, X1 , . . . , Xk 1 ).
Contractions have the following compatibility with the wedge product, similar to that
for the exterior differential:
iX (a ^ b ) = iX a ^ b + ( 1)k a ^ iX b , (6.8)
LX ( f ) = X( f ), LX (d f ) = dX( f ),
LX (a ^ b ) = LX a ^ b + a ^ LX b (6.9)
Proof. As in the case of the exterior differential, we can use the product rule to show
that LX is local: (LX a)|U depends only on a|U and X|U . Since any differential form
is a sum of wedge products of 1-forms, LX is uniquely determined by its action on
functions and differential of functions. This proves uniqueness. For existence, we
give the following formula:
LX = d iX + iX d.
LX f = iX d f = X( f ),
LX d f = diX d f = dLX f = 0.
t
u
6.8 Lie derivatives and contractions 119
Exercise 96. For each of the following vector-fields X and differential forms a
on R3 (with coordinates (x, y, z)) compute LX a
(a) X = y ∂∂x + x ∂∂y + z ∂∂z and a = ydx xdy zdz.
(b) The Wikipedia example: X = sin x ∂∂y y2 ∂∂x and a = x2 sin(y).
(c) The Wikipedia example: X = sin x ∂∂y y2 ∂∂x and a = (x2 + y2 )dx ^ dz.
These have the following compatibilities with the wedge product: For a 2 W k (M)
and b 2 W l (M) one has
One says that LX is an even derivation relative to the wedge product, whereas d, iX
are odd derivations. They also satisfy important relations among each other:
d d=0
LX LY LY LX = L[X,Y ]
iX iY + iY iX = 0
d LX LX d = 0
LX iY iY LX = i[X,Y ]
iX d + d iX = LX .
Again, the signs are determined by the even/odd parity of these operators; one should
think of the left hand side as ‘graded’ commutators, where a plus sign appears when-
ever two entries are odd. Writing [·, ·] for the graded commutators (with the agree-
ment that the commutator of two odd operators has a sign built in) the identities
become
[d, d] = 0,
[LX , LY ] = L[X,Y ] ,
[iX , iY ] = 0,
[d, LX ] = 0,
[LX , iY ] = i[X,Y ] ,
[d, iX ] = LX .
120 6 Differential forms
This collection of identities is referred to as the Cartan calculus, after Élie Cartan
(1861-1951), and in particular the last identity (which certainly is the most intrigu-
ing) is called the Cartan formula. Basic contributions to the theory of differential
forms were made by his son Henri Cartan (1906-1980), who also wrote a textbook
on the subject.
Exercise 98.
(a) As an illustration of the Cartan identities, let us prove the following formula
for the exterior differential of a 1-form a 2 W 1 (M):
Exercise 100. Prove the Jacobi-identity for the Lie derivative: for any X,Y, Z 2
M we have
L[X,[Y,Z]] + L[Y,[Z,X]] + L[Z,[X,Y ]] = 0.
6.8.1 Pull-backs
F ⇤ (b1 ^ b2 ) = F ⇤ b1 ^ F ⇤ b2 .
F ⇤ (du ^ dv).
Proof.
F ⇤ b = dF 1 ^ · · · ^ dF n
∂ F1 ∂ Fn
= Â · · · i dxi1 ^ · · · ^ dxin
i1 ...in ∂ x ∂x n
i 1
∂ F1 ∂ F n s(1)
= Â ∂ x s(1)
· · ·
∂ x s(n)
dx ^ · · · ^ dxs(n)
s2Sn
∂ F1 ∂ Fn
= Â sign(s) ∂ xs(1) · · · ∂ xs(n) dx1 ^ · · · ^ dxn
s2Sn
= J dx1 ^ · · · ^ dxn ,
Here we noted that the wedge product dxi1 ^ · · · ^ dxin is zero unless i1 , . . . , in are a
permutation of 1, . . . , n. t
u
One may regard this result as giving a new, ‘better’ definition of the Jacobian deter-
minant.
Remark 6.3. The Lie derivative LX a of a differential form with respect to a vector
field X has an important interpretation in terms of the flow Ft of X. Assuming for
simplicity that X is complete (so that Ft is a globally defined diffeomorphism), one
has the formula
d
LX a = F ⇤ a.
dt t=0 t
(If X is incomplete, the flow Ft is defined only locally, but the definition still works.)
To prove this identity, it suffices to check that the right hand side satisfies a product
rule with respect to the wedge product of forms, and that it takes on the correct values
on functions and on differentials of functions. The formula shows that LX measures
to what extent a is invariant under the flow of X.
This definition does not depend on the choice of oriented chart. Indeed, suppose
(V, y) is another oriented chart with supp(w) ✓ V , and write
1 ⇤
(y ) w = g dy1 ^ · · · ^ dym .
as required.
Remark 6.4. Here we used the change-of-variables formula from multivariable cal-
culus. It was very important that the charts are oriented, so that J > 0 everywhere.
Indeed, for general changes of variables, the change-of-variables formula involves
|J| rather than J itself.
If w is not necessarily supported in a single oriented chart, we proceed as follows.
Let (Ui , ji ), i = 1, . . . , r be a finite collection of oriented charts covering supp(w).
Together with U0 = M\supp(w) this is an open cover of M.
Lemma 6.5. Given a finite open cover of a manifold there exists a partition of unity
subordinate to the cover, i.e. functions ci 2 C• (M) with supp(ci ) ✓ Ui and Âri=0 ci =
1.
k The support of a form is defined similar to the support of a function, or support of a vec-
tor field. For any differential form a 2 W k (M), we define the support supp(a) to be the
smallest closed subset of M outside of which a is zero. (Equivalently, it is the closure of
the subset over which a is non-zero.)
6.11 Stokes’ theorem 123
Indeed, partitions of unity exists for any open cover, not only finite ones. A proof is
given in the appendix on ‘topology of manifolds’.
Let c0 , . . . , cr be a partition of unity subordinate to this cover. We define
Z r Z
w=Â ci w
M i=1 M
where the summands are defined as above, since ci w is supported in Ui for i 1. (We
didn’t include the term for i = 0, since c0 w = 0.) We have to check that this is well-
defined, independent of the choices. Thus, let (V j , y j ) for j = 1, . . . , s be another
collection of oriented coordinate charts covering supp(w), put V0 = M supp(w),
and let s0 , . . . , ss a corresponding partition of unity subordinate to the cover by the
Vi ’s.
Then the Ui \V j form an open cover, with the collection of ci s j as a partition of
unity. We obtain
s Z s Z r s r Z
 s jw =  (  ci ) s j w =  s j ci w.
j=1 M j=1 M i=1 j=1 i=1 M
R
This is the same as the corresponding expression for Âri=1 M ci w.
Of course, this definition works equally well for any smooth map from S into M. For
example, the integral of compactly supported 1-forms along arbitrary paths g : R !
M is defined. Note also that M itself does not have to be oriented, it suffices that S is
oriented.
D = {p 2 M| f (p) 0}.
124 6 Differential forms
∂ D = {p 2 M| f (p) = 0},
Example 6.1. The region with bounday defined by the function f 2 C• (R2 ), given
by f (x, y) = x2 + y2 1, is the unit disk D ✓ R2 ; its boundary is the unit circle.
Example 6.2. Recall that for 0 < r < R, zero is a regular value of the function on R3 ,
p
f (x, y, z) = z2 + ( x2 + y2 R)2 r2 .
The corresponding region with boundary D ✓ R3 is the solid torus, its boundary is
the torus.
(Indeed, given an oriented submanifold chart for which D lies on the side where
x1 0, one obtains a region chart by composing with the orientation-preserving co-
ordinate change (x1 , . . . , xm ) 7! ( x1 , x2 , x3 . . . , xm ).) We call oriented submanifold
charts of this kind ‘region charts’.††
Lemma 6.6. The restriction of the region charts to ∂ D form an oriented atlas for
∂ D.
Proof. Let (U, j) and (V, y) be two region charts, defining coordinates x1 , . . . , xm
and y1 , . . . , ym , and let F = y j 1 : j(U \ V ) ! y(U \ V ), x 7! y = F(x). It
restricts to a map
∂ y1 ∂ y1
> 0, = 0, for j > 0.
∂ x1 x1 =0 ∂xj x1 =0
⇤⇤ Note that while we originally defined submanifold charts in such a way that the last m
k coordinates are zero on S, here we require that the first coordinate be zero. It doesn’t
matter, since one can simply reorder coordinates, but works better for our description of
the ‘induced orientation’.
†† This is not a standard name.
6.11 Stokes’ theorem 125
Hence, the Jacobian matrix DF(x)|x1 =0 has a positive (1, 1) entry, and all other en-
tries in the first row equal to zero. Using expansion of the determinant across the first
row, it follows that
∂ y1
det(DF(0, x2 , . . . , xm )) = det(DF 0 (x2 , . . . , xm )).
∂ x1 x1 =0
where the ci are supported in Ui and satisfy Âi ci over D \ supp(w). By the same
argument as for D = M, this definition of the integral is independent of the choice
made.
Theorem 6.3 (Stokes’ theorem). Let M be an oriented manifold of dimension m,
and D ✓ M a region with smooth boundary ∂ D. Let a 2 W m 1 (M) be a form of
degree m 1, such that supp(a) \ D is compact. Then
Z Z
da = a.
D ∂D
R
∂ D i a, where i : ∂ D ,! M is the
As explained above, the right hand side means ⇤
inclusion map.
Proof. We will see that Stokes’ theorem is just a coordinate-free version of the fun-
damental theorem of calculus. Let (Ui , ji ) for i = 1, . . . , r be a finite collection of re-
gion charts covering supp(a) \ D. Let c1 , . . . , cr 2 C• (M) be functions with ci 0,
supp(ci ) ✓ Ui , and such that c1 + . . . + cr is equal to 1 on supp(a) \ D. (E.g., we
may take U1 , . . . ,Ur together with U0 = M\supp(w) as an open covering, and take
the c0 , . . . , cr 2 C• (M) to be a partition of unity subordinate to this cover.) Since
Z r Z Z r Z
da = Â d(ci a), a=Â ci a,
D i=1 D ∂D i=1 ∂ D
126 6 Differential forms
with compactly supported fi where the hat means that the corresponding factor is to
be omitted. Only the i = 1 term contributes to the integral over ∂ D = Rm 1 , and
Z Z
a= f1 (0, x2 , . . . , xm ) dx2 · · · dxm .
Rm 1
Let us integrate each summand over the region D given by x1 0. For i > 1, we have
Z • Z •Z 0
∂ fi
··· (x1 , . . . , xm )dx1 · · · dxm = 0
• • • ∂ xi
where we used Fubini’s theorem to carry out the xi -integration first, and applied the
fundamental theorem of calculus to the xi -integration (keeping the other variables
fixed, the integrand is the derivative of a compactly supported function). It remains
to consider the case i = 1. Here we have, again by applying the fundamental theorem
of calculus,
Z Z • Z •Z 0
∂ f1
da = ··· (x1 , . . . , xm )dx1 · · · dxm
D • • • ∂ x1
Z • Z • Z
= ··· fm (0, x2 , . . . , xm )dx2 · · · dxm = a
• • ∂D
t
u
As a special case, we have
Corollary 6.1. Let a 2 W m 1 (M) be a compactly supported form on the oriented
manifold M. Then Z
da = 0.
M
Note that it does not suffice that da has compact support. For example, if f (t) is a
function
R
with f (t) = 0 for t < 0 and f (t) = 1 for t > 0, then d f has compact support,
but R d f = 1.
A typical application of Stokes’ theorem shows that for a closed form w 2
W k (M), the integral of w over an oriented compact submanifold does not change
with smooth deformations of the submanifold.
6.11 Stokes’ theorem 127
Ft = F(t, ·) : S ! M.
does not change under deformations (isotopies) of the loop. In particular, g cannot
be deformed into a constant map, unless the integral is zero. The number
Z
1
w(g) = g ⇤w
2p S1
is the winding number of g. (One can show that this is always an integer, and that
two loops can be deformed into each other if and only if they have the same winding
number.)
Example 6.5 (Linking number). Let f , g : S1 ! R3 be two smooth maps whose im-
ages don’t intersect, that is, with f (z) 6= g(w) for all z, w 2 S1 (we regard S1 as the
unit circle in C). Define a new map
f (z) g(w)
F : S1 ⇥ S1 ! S2 , (z, w) 7! .
|| f (z) g(w)||
is called the linking number of f and g. (One can show that this is always an integer.)
Note that if it is possible to deform one of the loops, say f , into a constant loop
through loops that are always disjoint from g, then the linking number is zero. In his
case, we consider f , g as ‘unlinked’.
Example 6.6. Let M be a compact,R
connected oriented manifold. There exists a dif-
ferential form w on M such that M w = 1. (Note that w cannot be exact, since other-
wise the integral would be zero, by Stokes.) Given another compact, oriented mani-
fold N of the same dimension (e.g., M itself), and a smooth map F : N ! M, we can
define the degree of F
6.12 Volume forms 129
Z
deg(F) = F ⇤ w.
N
The degree is invariant under deformations of F. It turns out that it is independent
of the choice of w, and is always an integer. In the preceding example, the linking
number was defined as the degree of a map S1 ⇥ S1 ! S2 obtained from f , g.
Example 6.7. The Euclidean space Rn has a standard volume form G0 = dx1 ^ · · · ^
dxn . Suppose S ✓ Rn is a submanifold of dimension n 1, and X a vector field that
is nowhere tangent to S. Let i : S ! Rn be the inclusion. Then
G := i⇤ iX G0 2 W n 1 (S)
is a volume form. For instance, if S is given as a level set f 1 (0), where 0 is a regular
Example 6.8. Let i : Sn ! Rn+1 be the inclusion of the standard n-sphere. Let X =
Âni=0 xi ∂∂xi . Then
n
iX (dx0 ^ · · · ^ dxn ) = Â ( 1)i xi dx1 ^ · · · dxi 1
^ dxi+1 ^ · · · ^ dxn
i=0
Proof. We have to check that the condition is consistent. Suppose (U, j) and (V, y)
are two charts, where (j 1 )⇤G = f dx1 ^ · · · ^ dxm and (y 1 )⇤G = g dy1 ^ · · · ^ dym
with f > 0 and g > 0. If U \ V is non-empty, let F = y j 1 : j(U) ! y(V ) be
the transition function. Then
1 ⇤ 1 ⇤
F ⇤ (y ) G |U\V = (j ) G |U\V ,
hence
g(F(x)) J(x) dx1 ^ · · · ^ dxm = f (x) dx1 ^ · · · ^ dxm .
where J is that Jacobian determinant of the transition map F = y j 1 . Hence f =
J (g F) on j(U \V ). Since f > 0 and g > 0, it follows that J > 0. Hence the two
charts are oriented compatible. tu
G 0 = fG , f > 0.
is a volume form on Ua ; on overlaps Ua \Ub these are related by the Jacobian deter-
minants of the transition functions, which are strictly positive functions. Let {ca } be
a locally finite partition of unity subordinate to the cover {Ua }, see Appendix A.4.
The forms ca Ga have compact support in Ua , hence they extend by zero to global
forms on M (somewhat imprecisely, we use the same notation for this extension).
The sum
G = Â ca Ga 2 W m (M)
a
is a well-defined volume form. Indeed, near any point p at least one of the summands
is non-zero; and if other summands in this sum are non-zero, they differ by a positive
function.
For a compact manifold M with a given volume form G 2 W m (M), one can define
the volume of M, Z
vol(M) = G.
M
Here the orientation used in the definition of the integral is taken to be the orientation
given by G . Thus vol(M) > 0.
Note that volume forms are always closed, for degree reasons (since W m+1 (M) =
0). But on a compact manifold, they cannot be exact:
6.12 Volume forms 131
Theorem 6.6. Let M be a compact manifold with a volume form G 2 W m (M). Then
G cannot be exact.
R
Proof. We Rhave vol(M) = MG > 0. But if G were exact, then Stokes’ theorem
would give M G = 0.