Differential Forms and Applications (Do Carmo)
Differential Forms and Applications (Do Carmo)
Manfredo P. do Carmo
Differential Forms
and Applications
With 18 Figures
This is a transJation of the Portuguese book "Formas Diferenciais e Apli~6es", first published by
IMPA in 1971.
ISBN 978-3-540-57618-1
{(dxi)p; i = 1,2,3}
is in fact the dual basis of {( ei) p} since
( ) ( ) aXi
dXi p ej = aXj =
{O,1, if i # j
if i = j.
or
2 1. Differential Forms in R n
3
w = Laidxi,
i=1
where ai are real functions in R 3 • If the functions ai are differentiable, w is
called a differential form of degree 1.
Now let A2(R!)* be the set of maps <P: R! x R! - t R that are bilinear
(i.e., <p is linear in each variable) and alternate (i.e., <P~Vl.V2) = -<p(v2,vd).
With the usual operations of functions, the set A2(Rp)* becomes a vector
space.
When <P1 and <P2 belong to (R!)*, we can obtain an element <P1 1\ <P2 E
A2(R!)* by setting
(<p11\ <P2)(V1, V2) = det(<pi(vj))
The element (dXi)p 1\ (dxj)p E A2(R!)* will be denoted by (dXi 1\ dXj)p' It is
easy to see that the set {(dXi I\dxj)p, i < j} is a basis for A2(R!t (this will
be proved in a more general setting in Proposition 1 below). Furthermore,
and
<P: R; x ... x R; -t R
, I
"
k times
(alternating means that <P changes signs with the interchange of two con-
secutive arguments). With the usual operations, Ak(R;)* is a vector space.
Given <Pl. ... , <Pk E (R;)*, we can obtain an element <P1 1\ <P2 1\ ... 1\ <Pk of
Ak(R;)* by setting
1. Differential Forms in R n 3
It follows from the properties of determinants that CPl/\ CP2/\ ••• /\ CPk is in fact
k-linear and alternate. In particular (dXh )p/\(dXh)p/\, .. /\ (dXik) E Ak(R;)*,
il, i2,' .. , ik = 1, ... , n. We will denote this element by (dXil /\ dXi2 /\ ... /\
dXik)P'
is applied to
L ail ... ik dXil /\ ... /\ dXik (ejt, ... , ejk) == ajt ... jk == O.
il <... <ik
for all il, ... ,ik. It follows that I == g. Setting I(eip ... ,eik) == ail ... ik' we
obtain the above expression for f. 0
where ah ... i" are real functions in R n. When the ail ... ;" are differentiable
functions, w is called a differential k-form.
For notational convenience, we will denote by I the k-upla (i l , ... ,i k ),
i l < ... < ik, ij E {I, ... , n}, and will use the following notation for w:
w = La/dx/.
/
Example 1. In R4 we have the following types of exterior forms (where ai, aij,
etc., are real functions in R4):
O-forms, functions in R 4 ,
I-forms, a1dxl + a2dx2 + a3dx3 + a4dx4,
2-forms, al2dxl /\ dX2 + al3dxl /\ dX3 + al4dxl /\ dX4 + a23dx2 /\ dX3 +
a24dx2 /\ dX4 + a34dx3 /\ dX4,
3-forms, al23dxl /\dX2 /\dX3 +al24dxl /\dX2 /\dX4 +al34dxl /\dX3 /\dX4 +
a234dx2 /\ dX3 /\ dX4,
4-forms, a1234dxl /\ dX2 /\ dX3 /\ dX4.
From now on, we will restrict ourselves to differential k-forms and we will
call them simply k-forms.
We are going to define some operations on k-forms in Rn.
First, if wand <p are two k-forms:
Next, if w is a k-form and <p is an s-form, we can define their exterior product
w /\ <p, which is an (s + k)-form, as follows.
Definition 4. Let
This follows immediately from the definition and will be left as an exercise
(Exercise 3).
The exterior product of forms in R n has the following properties.
Remark 2. Although dXi A dXi = 0, it is not true that for any form w Aw = O.
For instance, if
then
w A w = 2XIX2dxI /\ dX2 A dX3 A dX4.
See however Exercise 4.
One of the most important features of differential forms is the way they
behave under differentiable maps. Let f: R n -+ R m be a differentiable map.
Then f induces a map J* that takes k-forms in R m into k-forms in R n and is
defined as follows. Let w be a k-form in Rm. By definition, J*w is the k-form
in R n given by
Herep ERn, VI, ... ,Vk E R;, and dfp:R; -+ R/(p) is the differential of the
map f at p. We set the convention that if 9 is a O-form,
J*(g)=goJ.
We are going to show that the operation J* on forms is equivalent to
"substitution of variables". Before that, we need some properties of J*.
Proof. The proofs are very simple. Let pERn and let VI, •.• , Vk E R;.
Then
(a) J*(w + CP)(P)(VI, ... , Vk) = (w + cp)U(p))(dfp(vd, .. ·, dfp(Vk)) =
U*W)(P)(VI"'" Vk) + U*cP)(p)(vI, ... , Vk) = U*w+ J*cp)(p)(vI, ... , Vk).
(b) J*(gw)(P)(vI, ... , Vk) = (gw)U(p))(dfp(Vl), .. " dfp(Vk)) = (g 0 J)(p) .
J*w(p)(Vl, ... , Vk) = J*g(p) . J*w(p)(Vl, ... , Vk).
(c) By omitting the indication of the point p, we obtain
Remark 3. We will show below (See Proposition 4) that (c) holds not only
for 1-forms but for k-forms as well.
1. Differential Forms in R n 7
Since
!*(dYi)(V) = dYi(df(v)) = d(Yi 0 f)(v) = d/;(v),
we have
rw = ~ a[(h(xI,"" xn), ... , fm(Xl,"" xn))dh /\ ... /\ d/;k'
[
Proof. By setting (YI,. .. , Ym) = (It (Xl!' .. , xn), . .. , fm(xl!' .. , Xn)) E Rm,
(Xl! ... ,Xn) ERn, W = "'r.-rardYr, cp = "'r.-J bJdYJ, we obtain
Proposition 5.
a) d(WI + W2) = dwl + dw2, where WI and W2 are k-forms
b) d(w /\ <p) = dw /\ <p+ (_l)kw /\ d<p, where W is a k-form and <p is an s-form
c) d(dw) = d2w = O.
d) d(f*w) = f*(dw), where W is a k-form in R m and f: R n --+ R m is a
differentiable map.
Proof.
(a) is straightforward.
(b) Let W = '£r ardxr, <p = '£J bJdxJ. Then
= L d(arbJ) /\ dXr /\ dXJ
rJ
= L bJdar /\ dxr /\ dXJ + L ardbJ /\ dxr /\ dXJ
rJ rJ
= dw /\ <p + (_l)k L ardxr /\ dbJ /\ dXJ
rJ
= dw /\ <p + (_l)kw /\ d<p.
(c) Let us first assume that w is a O-form, i.e., w is a function f: R n --+ R
that associates to each (Xl, ... ,Xn ) ERn the value f(xl> ... ,xn ) E R. Then
·
Smce -.!f!.L - -.!f!.L
{}x;{}Xj - {}Xj{}x; an
d dXi /\ dXj -- - dXj /\ dXi, ~. r),
-I- . we 0 b ' t h at
tam
Now, let <p = L:[ aldx[ be a k-form. By using the above, and the fact
that !* commutes with the exterior product, we obtain
= EdU*(aI»I\!*(dx[» = E!*(daI)I\!*(dxI)
[ I
EXERCISES
cp = xdx - ydy,
1/1 = zdx /\ dy + xdy /\ dz,
() = zdy.
Compute: cp /\ 1/1, () /\ cp /\ 1/1, dcp, d1/1, d(}.
6) Let f: U C R m -+ R n be a differentiable map. Assume that m < nand
let w be a k-form in Rn, with k > m. Show that j*w = O.
7) Let w be the 2-form in R 2n given by
v(el, ... ,e n ) = 1,
where {e;}, i = 1, ... , n, is the canonical basis of R n. Show that:
a) If Vi = E aijej, then
V(Vl, ... , Vn) = det(aij) = vol(Vl, ... , Vn).
(the form v is called the volume element of R n)
b) v = dXl/\"'/\ dxn.
10) (Hodge star operation). Given a k-form w in R n we will define an (n- k)-
form *w by setting
and extending it linearly, where il < ... < ik, jl < ... < jn-k,
(il, ... ,ik,jl, ... ,jn-k) is a permutation of (I,2, ... ,n), and a is 0 or
1 according to the permutation is even or odd, respectively. Show that:
a) If w = a12dxl 1\ dX2 + a13dXl 1\ dX3 + a23dx2 1\ dX3 is a 2-form in R 3 ,
then
*w = a12dx3 - a13dx2 + a23dxl,
b) If w = a1dxl + a2dx2 is a I-form in R2, then
c) **w = (_I)k(n-k)w.
11) (The divergence). A differentiable vector field v in R n may be consid-
ered as a differentiable map v: R n - t R n. We will define a function
div v: R n - t R (to be called the divergence of v) as follows:
(div v)(P) = trace(dv)p, p ERn,
where (dv)p: R; - t R; is the differential of vat p. Show that:
a) If v = L: aiei, where {eil is the canonical basis of Rn, then
d. ,",oai
lV v = L...., OXi .
Notice that 131 /\ •.. /\ 13k /\ CPi = CPI /\ ... /\ CPk /\ CPi = O. This implies
that
L bi7)f3I /\ ... /\ 13k /\ (37) = 0,
7)
and since 131 /\ ... /\ 13k /\ f3T/ are linearly independent, biT/ = O.
14 1. Differential Forms in R n
c) If 1,0 = 1,01/\ ... /\l,Ok is decomposable, the vectors VI, . .. ,Vk E R;,where
Vi +-+ l,Oi is the correspondence induced by the inner product ( , ) of
R n , are linearly independent, and the subspace of R n generated by
them does not depend on the representation of 1,0. This will be called
the subspace of 1,0.
d) If 1,0 = 1,01/\' .. /\ I,Ok is decomposable, the k-volume of the solid gener-
ated by VI,' .. ,Vk does not depend on the representation of 1,0. This
will be called the volume of 1,0.
e) If 1,0 = 1,01 /\ ••• /\ I,Ok is decomposable, define *1,0 as an element of
/\n-k(R;)* with the following properties:
i) the subspace of *1,0 is perpendicular to the subspace of 1,0.
ii) the volume of *1,0 is equal to the volume of 1,0.
iii) 1,0 /\ *1,0 is positive, i.e., its value in a positive basis ofR; is positive.
Prove: That *1,0 is well defined.
Hint: Let Vi +-+ l,Oi, i = 1, ... , k, as in (c) and let W be the subspace
of R; generated by the v~s. In W 1. consider an orthonormal basis
ek+1, ... ,en so that the basis VI, ... ,Vr , ek+1, ... ,en of R; is positive.
Define 1,0j E (R;)*, j = k + 1, ... ,n, by the correspondence 1,0j +-+ ej
mentioned in (c), and let). > 0 be the volume of W. Check that
).1,Ok+1. /\ ••. /\ I,On satisfies (i), (ii) and (iii).
f) Let VI, V2 be two vectors in R 3 and let 1,01 +-+ VI, 1,02 +-+ V2 be the one-
forms given by the correspondence in (c). Define the vector product
VI x V2 +-+ *(1,01 /\ 1,02) and describe geometrically the vector VI x V2.
g) A k-form w in R n is decomposable if w(P) is decomposable for each
pERn. Every k- form in R n is a linear combination of decomposable
k-forms of the type dXil /\ ... /\ dXik' Show that with the above defini-
tion, *(dXil /\ ... /\ dXik) gives the same expression as in Exercise 10.
16) (Poincare's lemma for I-forms). Let
!(x, y, z) = 10 1 (a(tx, ty, tz)x + b(tx, ty, tz)y + c(tx, ty, tz)z) dt.
Show that df = w (This means that the condition dw = 0, which is
satisfied when w = df, is in R3 also sufficient for the existence of an !
such that w = df. We have used R 3 for notational convenience but the
result holds for R n. Also, if w is only defined in an open set U eRn,
the result still holds in a neighborhood of each p E U).
Hint: Notice that dw = 0 implies that
ab aa ac aa ab ac
ax = ay' ax = az' az = ay
and use the identity
1. Differential Forms in R n 15
(I d
a(x, y, z) = Jo dt (a(tx, ty, tz)t)dt
where aI, a2 and a3 denote the partial derivatives of a(tx, ty, tz) relative
to the first, second and third argument, respectively.
17) We say that a differentiable vector field v defined in an open set U eRn
derives locally from a potential if for each p E U then exists a neighbor-
hood V :3 p, V C U, and a differentiable function g: V -+ R (to be called
the potential) such that v = grad g.
a) Let v be as above and let w be the I-form corresponding to v, i.e.,
w(u) = (v,u), for all u ERn. Show that v derives locally from a
potential if and only if dw = 0.
Hint: Use the local form of Exercise 16.
b) Show that v derives locally from a potential if and only if rot v = 0.
c) Let v be the vector field (electric attraction):
1
v(P) = - (x2 + y2 + z2)3/2 (x, y, z), P E R3 - {O,O,O}
°
k+I
Hint: Notice that dw = implies that
ab aa ac aa ab ac
ax = ay' ax = az' az = ay'
and apply Euler's relation
16 1. Differential Forms in R n
* This chapter is not used in the rest of the book and can be omitted on a first
reading.
18 2. Line Integrals
1w=-/w.
-e e
1 = 1 = lb
w df c*(df) = f(c(b)) - f(c(a)),
that is, 1 w depends only on the end points of c. It also follows that if w is
Proof. We have already shown that (1) :::} (2) :::} (3). That (3) :::} (2) is
immediate. It remains to show that (2) :::} (l).
1
Let us assume (2) and fix a point p E V. For each x E V, let c be a
piecewise differentiable curve joining p to x. Define f: V --. R by f(x) = w.
By (2), f is well defined. We claim that df = w, which will conclude the proof.
Since df = '" ~f dXi,
L....- uX'
we must show that if(x)
x,
= ai(x), i = 1, ... , n.
i '
Let ei = (0, ... ,0,1,0, ... ,0), where 1 is in i-th place, and consider the curve
Ci:t--.x+tei, tE(-£,£),
that joins x to x + tei and is contained in V for t small. Then
2. Line Integrals 19
~f (x) = t--+o
UXi
lim ~{J(x + tei) -
t
f(x)}
= 1Im-
· 1{1 w- 1} = li
w m-11 w
lit
t--+O t c+c, c t--+O t c,
= lim -
t--+O t 0
ai(s)ds = ai(O) = ai(X).
o
Example 1. Consider the form
wo = - 2y
x +y
2dx + 2 2dy
x
x +y
defined in U = R2 - {(O, On. A direct computation shows that dwo = O.
Rather than doing this computation, we prefer a geometric approach. Choose
a half-line L issuing from the origin 0 = {(O, On and consider polar coordi-
nates (p, (J) in R2 -L. Since x = pcos((J+(Jo), y = psin((J+(Jo), we obtain that
wo = d(J in R2 - Lj here (Jo is the angle from Ox to L. Since L is arbitrary,
dwo = d2(J = 0, hence wo is closed. In addition, this shows that the form wo
is locally (that is, in a neighborhood V of each point of U) the differential of
-+
a function (J that measures the positive angle of Op with L, p E V. We call (J
an angle function and wo the element of angle (with respect to the origin 0).
Although the form wo is closed and locally exact, it is not exact in R2 -
{O}. To see this, consider the unit circle c(t) = (cos t, sin t), t E [0,211"], around
the origin. Then c: [0,211"] -- U, and
1 c(t)
wo =1
0
2w c·Wo = 12w dt = 211".
0
The fact that the closed form Wo in Example 1 is locally exact is a general
fact. More precisely.
f(q) = {
J(jet)
w= 11
0
{a«(3(t»(x - Xo) + b«(3(t»(y - Yo) + e«(3(t»(z - zo)}dt
= 1
o
1 {(oa
ax oy
oa - zo) ) t + a } dt =
oa - Yo) + -(z
-(x - xo) + -(y
oz
= 11 { (:t(a({3(t)))) t + a} dt =
{I d
= Jo dt (a«(3(t»t)dt = a«(3(I» = a(q).
2. Line Integrals 21
of of
oy (q) = b(q), oz (q) = c(q).
1 1;
w= L
•
w = L[J;(ti+d - J;(ti)].
•
(1)
1
Now, if c: [0, I] -+ U is only continuous, such a partition still exists and
we define w by (1).
We must show that this is independent of the chosen subdivision. This is
a standard argument and we will present it here for completeness.
Given a partition P, a refinement of P is a new partition obtained from
P by adding new points to it. If we add a point t' E (ti' ti+1), we have that
c(t) E B j • Since
(Ji(ti+d - J;(t')] + (Ji(t') - !i(ti)] = [J;(ti+d - !i(ti)],
the value of the integral for this new partition does not change. It follows
that the integral remains the same under refinements of a given partition.
Now, given two distinct partitions, we can form a third partition, by adding
to the first all the points of the second. This is a common refinement, the
integral of which must be equal to the integrals of the first and of the second
partitions, which are therefore equal. This shows the required independence.
Theorem 1 also raises the following question. We know that a closed form
is locally exact. When is it globally exact? We know that we need some
restriction on the domain of the closed form, since by Example 1, the form
element of angle is closed in R2 - {o} but it is not exact there.
The answer to this question will depend on how the integral of a closed
form along a continuous curve varies when we deform the curve continuously.
To make this notion precise, we introduce the notion of homotopy.
22 2. Line Integrals
Theorem 2. Let w be a closed 1-)<-1171 ... ejined in an open set U eRn. Let
CO,Cl: [a,bJ -+ U be two continuous hutnotopic curves in U. Then
1w=l w.
Co Cl
(4)
But the sides of Rjk that are interior to R appear twice with opposite orien-
tations in the above sum. So the corresponding integrals cancel out, and we
have
2. Line Integrals 23
O!j+l,k
t = tj+ 1
t
/3jk Rjk /3j,Hl
t = tj
1
o a b 8
Fig. 2.1
o= 1 + J[ -1
co
w
f3b
w
Cl
w - [ w,
Jf3a
((5))
where /3a is the curve H(a, t) and /3b is the curve H(b, t). Since these curves
reduce to points, the corresponding integrals vanish, and we are left with
1w-j w
Co - Cl '
as we wished. o
We should observe that if Co and Cl are closed curves (i.e., co(a) = co(b)
and Cl (a) = Cl (b)) that are freely homotopic, then, although the curves /3a
and f3b do not reduce to points, they are equal. Therefore, by (5), we still have
the equality in (4). For future reference, we will sum up the above discussion
in a proposition.
1
Proposition 2. Let w be a closed 1-form defined in U, and let Co and Cl be
two closed curves freely homotopic in U. Then w =j w; in particular,
~ r () = n(FjD)
211" lOD
the index of F in D. We claim that n(Fj D) is an integer.
To see that, let u = f(x,y), v = g(x,y), and let Wo = u:rt~~U be the
element of angle in the plane (u, v) relative to the origin (0,0). Then () = F*wo
and
n(Fj D) = -2 1
11"
r
(J = -2
lc
1
11"
r
F*wo = 21
lc
wo,
11"
r
lFoC
that is, n(Fj D) is the winding number (see Remark 1) of the closed curve
Foe around the origin in the plane (u, v). This shows that n( Fj D) is an
integer.
A consequence of the invariance under free homotopy of Jc () is the fol-
lowing existence theorem for solutions of the equation F = o.
Proposition 4. If n(Fj D) "I- 0, there exists some point qED such that
F(q) = O.
2. Line Integrals 25
Proof. Assume that no such q exists, and let p be the center of the disk D.
Then the map H: [0,211"] x [0,1] --* R2 - {o} given by
° °
It remains to show that there exists c > so that H(q, t) f:. for all q f:. 0,
Iql < c, and all t E [0, 1J. To do this, set G = 1/ IT-II and observe that, for
all q,
Iql = IT-ITql ~ IT-IIITql = l/G ITql ,
hence ITql ~ G Iql. Now, take c > 0 so that in the disk D of radius c we have
IR(q)1 ~ G/2 (recall that lim R(q) = 0). Then, if q f:. 0,
q--+O
Fig. 2.2
sector contains exactly one zero (Fig. 2.2). Let C i be a circle around Pi E Si
so small that is contained in Si. There is an obvious free homotopy taking
Ci in aSi. Therefore n(Fj D) is given by
as we wished to prove. o
EXERCISES
w = x 2 eX
+y
2 {(x cos y + y sin y)dy + (xsiny - ycosy)dx}
defined in R2 - {o}.
a) Show that w can be written as
w = eX cos y Wo + eX sin y d(log r),
J
where Wo is the element of angle at 0 and r = x 2 + y2 j check by
computation that dw = O.
b) Show that w - Wo satisfies the condition of Exercise 2(c), hence is
exact.
28 2. Line Integrals
1
4) Let w be a I-form defined in an open set U eRn. Assume that for each
closed differentiable curve c in U, w is a rational number. Prove that
w is closed.
5) Let U, VeRn be simply connected open sets such that Un V is con-
nected. Let w be a closed I-form so that w is exact in U and w is exact
in V. Show that w is exact in U U V.
6) (Applications to complex junctions). Line integrals are quite useful in
the study of complex functions J: C -+ C. Here the complex plane C
is identified with R2 by setting z = x + iy, z E C, (x,y) E R2. It is
convenient to introduce the complex differential form dz = dx + idy and
to write
J{z) = u{x, y) + iv(x, y) = u + iv.
Then the complex form
Assume that u and v are Cl. Recall that J is holomorphic if and only if
(Cauchy-Riemann equations.)
Show that:
a) J is holomorphic if and only if the real and imaginary parts of J{z)dz
are closed.
1
b) (Cauchy's theorem). If J is holomorphic in a simply-connected domain
U c C and c is a closed curve in U, then J{z)dz = O.
c) If J is holomorphic, the function J'(z) (the derivative of J in z) given
by the equation dJ d~f du + idv = J'(z)dz is well defined and J'(z) =
Ux - iu ll •
d) If J is holomorphic in U c C and J'{z) '# 0, z E U, then all zeroes of
J are simple and positive; furthermore, if D c U is a disk such that
there are no zeroes in aD, then
# of zeroes of J in D 1. [
= -2 dlf .
7I'Z laD J
Hint. That the zeroes of J are simple and positive follows from
2. Line Integrals 29
1
-.
2n
1 -Idf J
IJD
= -
1
271"
udv - vdu
u v
2 2 =
.
# of zeroes of 1 III D.
~ f
271" JIJDl
w = +1 -1
271"
18D2
w--l
- ,
Y C
Fig. 2.3
1
points Pl,P2 E U,
v = f(P2) - f(PI)
Hint. We use the notation of the hint in (a). O(J) is still constant
along the trajectories of v. Thus d(O(J)) = gw or dO· gw = gw, where
9 is a new integrating factor. Since w :f; 0, dO· 9 = g.
3. Differentiable Manifolds
Differential forms were introduced in the first chapter as objects in Rn; how-
ever, they, as everything else that refers to differentiability, live naturally in
a differentiable manifold, a concept that we will develop presently.
We will start with the most familiar example of a differentiable manifold,
namely a regular surface in R 3 . In what follows, we will use as a reference
"M. do Carmo, Differential Geometry of Curves and Surfaces, Prentice-Hall,
1976", to be quoted as [dC]. Let us recall (d. [dC), Chap. 2 §2.2) that a
subset S c R 3 is a regular surface if, for each pES, there exist a neighbor-
hood V of pin R3 and a map 10;: Uo; C R2 - t V n S of an open set Uo; in R2
onto V n S such that:
1) 10; is a differentiable homeomorphism.
2) the differential (dlo;)q: Tq(Uo;) - t R3 is injective for each q E Uo;.
The map 10;: Uo; - t S is called a parametrization of S around p.
The most important consequence of the above definition is the fact that
the change of parameters is a diffeomorphism. More precisely, if 10;: Uo; - t S
and 1{3: U{3 - t S are two parametrizations such that 10;(U0;) n 1{3(U{3) = W =I-
¢, then (cf. [dC], Theorem of §2.5) the maps
IiI o/o;:/;;I(W) -t R2,
1;;10 1{3: lil(W) -t R2
are differentiable. It follows that on a regular surface it makes sense to talk
about differentiable functions and to apply the methods of Differential Cal-
culus.
The most serious problem with the above definition is its dependence on
R3. Indeed the natural idea of an abstract surface is that of an object that
is, in a certain sense, two-dimensional, and to which we can apply, locally,
the Differential Calculus in R2. Such an idea of an abstract surface (Le.,
with no reference to an ambient space) has been foreseen since Gauss. It
took, however, about a century for the definition to reach the definitive form
that we present below. One of the reasons for this delay was that, even for
surfaces in R 3 , the fundamental role of the change of parameters was not
clearly understood (see Remark 1).
Since nothing is gained in limiting ourselves to dimension two, we will
present the definition for dimension n.
34 3. Differentiable Manifolds
2) For each pair a,/3, with fo.(Uo.) n f/3(U/3) = W ::I </>, the sets f~l(W)
and l;l(W) are open sets in R n and the maps f;l 0 10., f~l 0 f/3 are
differentiable (Fig. 3.1).
-
Fig. 3.1
Example 1. The real projective plane p2(R). We will denote by P2(R) the
set of all lines in R3 that pass through the origin (0,0,0) of R 3 , i.e., p2(R)
is the set of "directions" in R 3 .
We want to introduce a differentiable structure in p2(R). For that, let
(x, y, z) E R 3, and notice that p2(R) is the quotient space of R3 - {(O, 0, O)}
by the equivalence relation rv:
Example 1 '. (The real projective space). Example 1 can easily be generalized.
Let (Xl. . .. ,xn+t> E R n +1 and define the n-dimensional real projective space
pn(R) as the quotient space of R n +1 - {OJ by the equivalence relation:
Example 2. (The Klein bottle). The Klein bottle is the subset of R4 defined
as follows (Fig. 3.2). Let Ox, Oy, Oz, Ow be the four coordinate axis in R4.
Let S be a circle with radius r, contained in the plane xOz, with center C
in the axis Ox so that C is at a distance a > r from O. The Klein bottle
is generated by rotating this circle around 0 z in such a way that when the
center C has described a rotation of an angle u in the plane xOy, the plane of
S has described a rotation of angle u/2 around OC in the 3-space OCOzOw
(this is possible because we are in R 4 ).
Let u and v be as in Fig. 3.1, and let Ul C R2 be given by
Ul = {(u,v) E R2;0 < u < 27r,0 < V < 27r}
3. Differentiable Manifolds 37
Fig. 3.2
It is clear that /1 (Ud contains the points ofthe Klein bottle that are not
on the circles u = 0 and v = O. We claim that /1 is injective.
To see that, let us first assume that z =I O. Then both sin v and cos u/2
are nonzero. Since 0 < u/2 < 7T, the expression -; = tan ~ determines u. By
knowing u, we can use that
. w +v'X2 +y2 - a
smv= -'-u' cos v = -~---"'--
rsmI r
is injective. Notice that b(U1 ) n !2(U2) = W is not connected but has two
connected components:
7r 7r
WI = {b(u,v); 2" < u < 27r}, W2 = {b(u,v);O < u < 2"}.
The change of coordinates is given by:
'it - u - 7r/2
1:;10 b { _-
v= v
(J • • (.~
Torus
B
0
A
A
0
B - 0J (]
~
Klein bottle
Fig. 3.3
In particular, it follows from the above that we can talk about differen-
tiable functions (<p: M n - t R) and differentiable curves (<p: 1 c R - t Mn) on
a differentiable manifold (1 c R will always denote an open interval of the
real line containing the origin 0 E R).
d (
dt <p 0 0
)\
t=O =~
i=l
a<p dXi \
a
(~ '( ) a )
L..J ax- dt t=O = L..J Xi 0 ax-
i=l a
<po
40 3. Differentiable Manifolds
respectively. Thus
Proof. We have just shown that TpM C T,. Conversely, if VET" then
v = Ei Ai (Ix;)o' Let a: I -+ M be given in the parametrization I by
Xi = Ait. Then a'(O) = v, i.e., v E TpM. 0
3. Differentiable Manifolds 41
With the notion of a tangent space we can define the notions of differential
of a map tp: Mf -4 Mr'
The linear map dtpp may be thought as a first order approximation of the
map t.p around p. Probably the most important local theorem in Calculus is the
inverse function theorem which states that if dtpp is an isomorphism then tp
is a local diffeomorphism at p. Being a local theorem, it extends immediately
to differentiable manifolds.
W = '~Yi
" 0. 8
8x~'
i •
Define a map Fa: Uo. X R n --+ TM by
42 3. Differentiable Manifolds
that is,
Example 4.
a) The curve a: R --+ R2 given by a(t) = (t 3 , t 2 ) is a differentiable map
but not an immersion. In fact, the condition of immersion is this case is
equivalent to the fact that a'(t) :f 0, which does not hold for t = 0.
b) The curve a(t) = (t 3 -4t, t 2 -4) is an immersion that is not an embedding,
since it has a self-intersection at (0,0) (for t = 2, t = -2).
c) The curve (Fig. 3.4)
=(0,-(t+2», tE(-3,-I)
a(t) { = a regular curve as in Fig. 3.2, t E (-1, -1/11")
= (-t,-sin(l/t», t E (-1/11",0)
is an immersion a: (-3,0) --+ R2 without self-intersections. However, a
is not an embedding. For, in the topology of R 2 , a neighborhood of a
point p in the vertical part of the curve has infinitely many connected
components, where as in the topology induced by a it is a connected
interval.
3. Differentiable Manifolds 43
Fig. 3.4
Example 6. It can be shown that the maps /;, i = 1,2,3, given in Example 2,
have injective differentials. This shows that the Klein bottle is a 2-surface of
R4.
where
Ul = {(Xl,X2, X3) E R3; Xl = 0, X~ + X~ < 1}
fi(X2,X3) = (D ll X2,X3), Dl = V~1---(-X-~-+-X-~-)
fl(x2,x3) = (-D ll X2,X3);
U2 = {(XllX2,X3) E R 3 ;X2 = O,X~ +X~ < 1}
ii(XllX3) = (xllD2,X3), D2 = J~1---(-x-i-+-X-5-)
Fig. 3.5
Let 7r: 8 2 - p2(R) be the canonical projection, that is, 7r(p) = {p, -p}
and notice that 7r(Ji+(Ui» = 7r(Ji-(Ui». We will define gi: Ui - P2(R) by
gi = 7r 0 it
Since the restriction of 7r to it(Ui) is injective, we obtain that
It follows that gil ogj is differentiable for all i,j = 1,2,3, hence {(Ui, gi)}
is a differentiable structure for p2(R).
3. Differentiable Manifolds 45
Actually the above structure and that of Example 1 will determine the
same maximal structure. For the coordinate neighborhoods are the same and
the coordinate changes are given (for i = 1, say) by
which is clearly differentiable. Notice also that it follows from the above that
the canonical projection 11": 8 2 - t p2(R} is a local diffeomorphism at each
point of 8 2 •
A A
AI---~~---IA
Fig. 3.6
From the above, it follows that a differential k-form in Mn is the choice, for
each parametrization (UOt, fOt) of M, of a differential k-form WOt in Uot in such
a way that for another parametrization (U{3, f(3), with fOt(UOt) n f{3(U{3) "I 1;,
we have Wot = (fi l 0 fOt)*w{3.
It is an important fact that all the operations defined for differential
forms in R n can be extend to differential forms in Mn through their local
representations. For instance, if W is a differential form in M, dw is the
differential form in M whose local representation is dw Ot . Since
Prool. We first prove that if such a Z exists, then it is unique. For that, let
I: U --. M be a parametrization, and let
hence
(XY - Y X)cp =~
~ (~( ab 0 aa 0) a)
aXi - bi-aXi' -aXj cpo
~ ai-'
) ,o 0
Definition 10. The vector field determined by the above lemma is called
the bracket [X, Y] = XY - YX of X and Y, and it is clearly differentiable.
Prool. (a) and (b) are immediate. To prove (c), we observe that
[[X,Y],Z] = [XV - YX,Z] =XYZ - YXZ- ZXY +ZYX
= [X, [Y, ZII + [V, [Z, XII,
3. Differentiable Manifolds 49
and use (a) to obtain (c). The proof of (d) is a direct and simple computation.
o
There exists an interesting relation between exterior differentiation of dif-
ferential forms and the bracket operation. For the case of I-forms, this relation
is as follows.
Notice that if (2') holds for WI and W2, it also holds for WI + W2. Thus, it
suffices to show that
50 3. Differentiable Manifolds
i=l
We will conclude this chapter with the global notion of orientability for
manifolds.
EXERCISES
1) Show with details that the real projective space pn(R) is a differentiable
manifold.
3. Differentiable Manifolds 51
x 2 - y2 = a, xy = b, xz = c, yz = d. (*)
It suffices to check that, under the condition x 2 + y2 + Z2 = 1, the
above equations have only two solutions which are of the form (x, y, z)
and (-x, -y, -z). The three last equations give:
If b, c, d are all zero, the equation (*) shows that at least two of the
coordinates x, y, z are zero, the remaining one being ±1, since x 2 +
y2 + Z2 = 1. If one of the values b, c, d is non zero the equation (**)
together with x 2+y2 + z2 = 1 will determine x 2, y2, z2 . From equations
(*), we see that the choice of a sign for one of the x, y, z determines
the signs of the remaining two.
6) Consider the cylinder C = {(x,y,z) E R3;x 2 +y2 = I} and identify the
point (x, y, z) with (-x, -y, -z). Show that the quotient space of C by
this equivalence relation can be given a differentiable structure (infinite
Mobius band).
7) Show that the tangent bundle of a differentiable manifold is orientable
(even if the manifold is not so).
8) Let M be a differentiable manifold that can be covered by two coordinate
neighborhoods VI and V2 in such a way that the intersection VI n V2 is
connected. Show that M is orientable.
9) Show that the sphere 8 n = {p ERn+!; Ipi = I} is orientable.
52 3. Differentiable Manifolds
10) Show that the real projective plane p 2 (R) is not orientable.
Hint. Show that if a manifold M is orientable, any open set in M is an
orient able manifold. Notice that p2(R) contains an open Mobius band
that is not orient able (cf. [de], §2.6, Example 3).
11) Show that the Klein bottle is not orientable.
12) A field of planes in an open set U C R3 is a correspondence P that
associates to each p E U a plane P(p) passing through p. The field P is
differentiable if the coefficients of the equation of P(p) are differentiable
functions in p. A integral surface of P is a surface S c R3 such that for
each q E S, we have TqS = P(q), i.e., S is tangent at each of its points
to the plane of the field passing there.
Let W be a differentiable I-form in U C R3 with w(q) =F 0, q E U. Show
that:
(a) W determines a differentiable plane field P by the condition
v E P(p) C R3 ~ wp(v) = 0.
(b) If S is an integral surface of P passing through p, then i*(w) = 0,
where i: S C R3 is the inclusio' .
(c) If there exists an integral' iacf' J of P passing through p, for all
p E U, there exists a I-form (f ill a neighborhood V C U of p such
that dw = w /I. (1.
Hint for (c): Consider two I-forms W2, W3 in such that w = WI, W2, W3 are
linearly independent in V, and write
dw = aW2 /I. W3 + {3w3 /I. WI + ')'WI /I. W2
°
By using the fact that d(i*w) = i*dw = and that the 2-form W2 /I. W3 is
nonzero, one concludes that a = 0, hence
15) Consider in the real line R the two following differentiable structures: 1)
(R,fd where fI(x) = x. 2) (R,h), where h(x) = x 3 • Show that:
a) The identity map i: (R, ft} -+ (R, h), i(x) = x, is not a diffeomor-
phism (thus the maximal structures determined by (R, ft} and (R, h)
are distinct)
b) The map t.p: (R, fI) -+ (R, h) given by t.p(x) = x 3 is a diffeomorphism
(thus, although the differentiable structures are distinct, they define
differentiable manifolds that are diffeomorphic).
16) (The orientable double covering). Let M be a connected differentiable
manifold. For each p E M, denote by Op the quotient space of the set
of all bases of TpM under the following equivalence relation: two bases
are equivalent if they are related by a matrix with positive determinant.
Clearly Op has two elements, and each element Op of Op is called an
orientation at p. Now let
where (XI, ... ,Xn ) E Ua , and [/xt, ... ,kl denotes the element ofOp
determined by this basis. Show that:
a) If {(Ua , fa)} is a differentiable structure in M, then {(Ua , j~)} is a
differentiable structure in if which is orient able (even if M is not).
b) The map 7r: if -+ M given by 7r(p,Op) = P is differentiable, sur-
jective, and each point p E M has a neighborhood V whose inverse
image 7r- 1 (V) is the disjoint union of two open sets each of which is
applied by 7r diffeomorphically onto V. By this reason, if is called
the orientable double covering of M.
c) Mis orient able if and only if if is not connected.
Hint. Notice that if M is orientable, the sets Ml = {(p, orient. of M in
pH, M2 = {(p, orient. opposite to that of M in p)}, are nonempty, dis-
joint, open subsets of if. For the converse, one has to show first that the
image 7r(F) of a closed set F c if is a closed set in M; this follows from
the fact that for each p E M, the inverse image 7r- I (p) has two points.
Now, if if is not connected, denote by C a connected component of if.
Then 7r(C) is an open, closed and nonvoid subset of M, hence 7r(C) = M.
Therefore, if is the disjoint union of two connected components and 7l"
applies diffeomorphically each such component onto M. It follows that
M is orient able.
17) (A non-Hausdorff manifold). Let S be the set given by the disjoint union
of R2 with a point p*. Let fI and 12 maps of R2 in S defined by
54 3. Differentiable Manifolds
we define
1 1 J
M
w=
v",
Wa =
U",
aadXl, ... ,dxn,
For that, we can assume, by contracting UOl and U/3 if necessary, that
VOl = V/3' Let the change of coordinates
1 = 1;1 01/3: U/3 -+ UOl
be given by
Xi = J;(Yl,"" Yn), i = 1, ... , n,
(XlI""X n) E UOl , (Yl, ... ,Yn) E U/3'
Since w/3 = /*(wOl ), we obtain that
w/3 = det(df)a/3dYI 1\ •.• 1\ dYn>
where
a/3 =aOl (/l(Yl, .. ·,Yn),···,ln(Yl,··.,Yn))'
On the other hand, by the formula of change of variables for multiple integrals
in R n, we obtain that
a) LCPi = 1,
i=1
b) 0:5 CPi :5 1, and the support of CPi is contained in some VOl; = Vi.
The family {CPi} is called a differentiable partition 01 unity subordinate to
the covering {VOl}' (When M is orientable, we choose {VOl} compatible with
the orientation).
Let us assume, for the time being, the existence of such a family. Assume
furthermore that M is orientable. We will define the integral of an n-form W
on M n as follows.
The support of the form CPiW is contained in Vi. By a previous definition,
it makes sense to write
1. Integration of Differential Forms 57
[
1M
W = fi=11M[ CPiW ,
the only question being whether this definition is independent of the choices
made.
Consider another covering {W.a} of M which determines on M the same
orientation as {Va}, and let {""i}, j = 1, ... ,s, be a partition of unity sub-
ordinate to {W.a}. Then {Va n W.a} will be a covering for M and the family
CPi""i will be a partition of unity subordinate to {Va n W.a}. Thus
where in the last equality it was used that, for each i, the functions CPi""i are
defined in Vi. Similarly,
~ 1, if tE(-2,-I),
(3(t) = 1, if t~-l.
- 1
-2 o -2 -1 o
lea) /l(b)
l(c)
°
Proof. Let (x~, ... ,x~) E U be such that g(x~, ... ,x~) = p. Since U is open,
there exists an r > such that Br(x~, ... , x~) c U. Let T the translation
in R n that takes (x~, ... , x~) to (0, ... ,0), and let H: R n -+ R n be the map
1. Integration of Differential Forms 59
that to each p E R n associates the point ~p. Then HoT takes Br(xY, . .. ,x~)
to B3(O).
a) L'Pi =1
i=l
b) o $ 'Pi $ 1, and the support of 'Pi is contained in some Val of the covering
{Va}.
2. Stokes Theorem
In this section we intend to establish Stokes theorem. For that, we will need a
series of definitions that will make it possible to state precisely the theorem;
once stated, the proof of the theorem is relatively simple.
For the two-dimensional case, a rough description of the theorem is as
follows.
Let w be a differential I-form defined on a two-dimensional, oriented,
manifold M2, and let dw be its exterior differential. Consider a region R of
M2 bounded by a closed regular curve C = oR. The orientation of R induces
an orientation on C, and the inclusion i: C ---. M allows us to consider the
restriction i*w of w to C. Under these conditions, Stokes theorem states that
the integral of the 2-form dw in R is equal to the integral of i*w in oR = C.
So, in a certain sense, the operators d (applied to forms) and 0 (applied to
smooth domains) are dual to each other.
If, in particular, M2 = R2 and w = Pdx + Qdy, Stokes theorem reduces
to Gauss theorem:
We now start to present the definitions we need and which are useful in
other contexts. The first one is an extension of the notion of a manifold to
include manifolds with "boundary". The definition of a manifold does not
include, for instance, the set M, given by
M = {(x, y, z) E R 3 ; Z = x 2 + y2, z:::; Zo, Zo > O}
(M is the closed set of the rotation paraboloid bounded above by Z = zo),
because the intersection VnM of any neighborhood V of a point p = (x, y, zoJ
in the "boundary" of M with M is not homeomorphic to an open set of R
(Fig. 4.2). Notice, however, that V n M is homeomorphic to an open set of
the closed half-space {(Xl,X2) E R2; Xl :::; OJ, whereas points of M that are
not in the boundary behave as points in a 2-manifold. This suggests a new
definition that will include the above situation.
A hall-space 01 R n is the set
H n = {(xt. ... ,Xn) ERn; Xl :::; OJ.
An open set of Hn is the intersection with Hn of an open U of R n.
We say that a function I: V ---. R defined in an open set V of Hn is
differentiable if there exists an open set U :::> V and a differentiable function
7 in U such that the restriction of 7 to V is equal to I. In this case, the
differential dIp, p E V, of I at p is defined to be dIp = d7 p.
When V does not contain points of the form (0, X2, ... , Xn), V is an open
set of R n and the definition of dIp agrees with the usual one. If p is of the form
2. Stokes Theorem 61
-
(X,1I,Zo)
-11
Fig. 4.2
(0,X2,'" ,xn ), dIp is defined for all tangent vector of curves in U passing
through p, i.e., for all vectors in R n with origin in p. Using such curves it is
easy to show that the definition of dIp is independent of the extension 1 of I.
In a similar way, we define a differentiable map I: V -+ R n •
'"
q
q2
;---
11 012
Fig. 4.3
III 012: U - t H n
such that the determinant of d(j-l oh)ql is nonzero. By the inverse function
theorem, 11-1 012 will take a neighborhood V C U of Q2 diffeomorphically
onto III 0 h(V). But then f1l 0 h(v) would contain points of the form
(Xl, ... , xn) with Xl > 0 which are not in Hn. This yields a contradiction
and completes the proof. 0
Let
2. Stokes Theorem 63
By identifying the set {(Xl. ... ,Xn) E Rnj Xl = O} with Rn-l, we see that
Ua is an open set in R n - 1 . By denoting by lathe restriction of fa to U a,
we see, by Lemma 3, that 101 (U a) C 8M. Finally, by letting p run in the
points of 8M, we easily check that the family {( U a,]a)} is a differentiable
structure for 8M. This proves the first part of the Proposition.
To prove the second part, assume that M is orient able and choose an
orientation for M, i.e., a differentiable structure {(Ua , fa)} such that the
changes of coordinates have positive jacobian. Consider the elements of
the family that satisfy the condition fa(Ua) n 8M i: ¢. Then the family
{(U a,fa)} described in the first part is a differentiable structure for 8M. We
want to show that if 1a (U a) n 1{3 (U {3) i: ¢, the change of coordinates has
positive jacobian, i.e., that
--1 -
det(d(J a 0 f {3)q) > 0,
for all q whose image, by some parametrization, is in the boundary.
Observe that the change of coordinates fa 0 til takes a point of the form
(0, xg, ... , x~) into a point of the form (0, x~, ... , x~). Thus, for a point q
whose image is in the boundary,
1 8x~ --1-
det(d(J;: 0 f{3) = -(J det(d(J a 0 f {3))'
8X1
r
JaM
i*w = r dw.
JM
Proof. Let K be the support of w. We will consider the following cases:
A) K is contained in some coordinate neighborhood V = f(U) of a
parametrization f: U c Hn --t M. In U,
n
W= L ajdX1 /\ ... /\ dXj_1 /\ dXj+l/\"'/\ dXn,
j=l
AI) Assume first that f(U)noM = ¢J. Then w is zero in oM and i·w = O.
Thus
f i·w = O.
IBM
We will show that
Fig. 4.4
2, Stokes Theorem 65
Then
1("{, )F'-I x,
u,
~ -1 aaj ) d Xl ' .. dX n = "
-a, ~( -1 ) '
,.
'-Ii
Q
aaj,dXl'" dX n
-a
x,
'''-If
,= ~(-IF [aj{xt. ... ,Xj-I,Xj,Xj+I,
0 ... Xn)
- aj{xt. ... ,Xj-l, x}, Xj+t. ..• ,Xn)]dXI ... dXj-1 dXj+1 ••. dXn = 0,
since aj(X1, .. ' ,x~, ... ,xn) = aj(xI,'" ,x}, ... ,xn ) = 0, for all j.
A2 ) Assume now that f{U) n aM :F ¢. Then the inclusion map i can be
written as: Xl = 0, Xj = Xj' Thus, using the induced orientation on the
boundary,
i*w = a1(0, X2,.'" Xn)dX2 1\ •.• 1\ dxn.
As in case (AI)' we will extend the functions aj to Hn, and will consider the
parallelepiped Q given by
j = 2, ... ,n
and such that the union of the interior of Q with the hyperplane
contains f-I(K), Then
Xl = °
1 dw = ~()'
~ -1'- 11 aaj
-.dXI···dxn
Q ax,
k
M j=l
+~
j=2
.11
L...,,( -1)'-
Q
0
[aj(Xl,"" Xj"", I
Xn) - aj(XI,"', Xj"", Xn)]
Since aj(xt. . .. ,x~,. '. ,Xn ) = aj(xt., " ,x}" " ,Xn ) = 0, for j = 2, ... ,n,
and al(xi,x2,"" xn) = 0, we obtain
Therefore,
o
Example. Let M be a bounded region of R3 such that the boundary 8M
of M is a regular hypersurface of R 3 j M is then a compact 3-dimensional
manifold with boundary 8M. Let v be a differentiable vector field in R 3 , and
let w be the I-form in R3 dual to v in the natural inner product of R3. Then
(cf. Exercise 11, Chap. 1) d(*w) = (div v)v, where v is the volume element
ofR3 •
Now choose an orientation for R3 and let N be the unit normal vector of
8M in the induced orientation. Finally, let a be the area element of 8M.
Consider, in a neighborhood U c R 3 of p EM, differentiable orthonormal
fields el, e2, N such that, in the points of 8M, el and e2 are tangent to 8M.
Then
i- * w(eb e2) = w(N) =< v, N >,
i.e., i-( *w) = (v, N)a. Thus, in this case, Stokes theorem
can be written as
[ div v v = [ < v, N > a
1M IBM
which is the well known divergence theorem in Analysis.
3. Poincare's Lemma
1 =1
e
w
e
dg ={
Joe 9 = 0,
Ie
and this contradicts the fact, easily computable, that W = 211'. It is possible,
however, to show that for each p E U there exists a neighborhood V C U of
p and a differentiable function gv in V such that dgv = w.
In this section, we will show that the situation of this example is com-
pletely general, that is, that the condition dw = 0 is a sufficient condition for
w to be locally exact (Those who are familiar with the material of Chapter 2
will notice that we are generalizing Theorem 1 ofthat Chapter). Actually, we
will prove the result in a form slightly more general which is more convenient
for the applications.
Proof. Let 11': M x R ---+ M be the projection 1I'(p, t) = p, and let w be the
k-form on M x R given by w = H*w, where H is the map given in the
definition of contract ability. We will need the following lemma.
w= Wl + dt 1\ "I, (1)
Now let it : M --+ M x R the map given by it(P) = (p, t); it is the inclusion
of Minto M x R at the "level" t. We will define a map I that takes k-forms of
M x R into {k -I)-forms of M as follows: If p EM and VI! V2, ••• , Vk E TpM,
then at p,
Therefore,
and
which completes the Case (b), and the proof of the lemma. o
Now we can complete the proof of the theorem (notice that so far we have
not used that dw = 0). Since M is contractible,
H 0 il = identity, H 0 io = const. = Po E M.
Thus
w = (H 0 iI)*w = ii{H*w) = i~w,
0= (H 0 io)*w= io(H*w) = iow.
Now, since dw = 0, we obtain that dW = H*dw = O. It follows by Lemma 5
that
w = iiw = d(Iw) = do:,
where 0: = lW. o
EXERCISES
M2 = ((x,y,z) E R3if(x,y,z) = a}
70 4. Integration on Manifolds; Stokes Theorem and Poincare's Lemma
is a regular orient able surface in R3 (cf. [de], §2.2). Show that the
restriction of W to M2 is the element of area of M2.
Hint: We want to show that if {VbV2} is a positive basis of Tp(M),
p E M then W(Vb V2) = area of the parallelogram made up by VI and V2.
Choose a parametrization 9( u, v) of M around p, compatible with the
orientation, and notice that
where (9u /l.9v)i is the i-th coordinate ofthe vector product 9u /l.9v in the
canonical basis ofR3. Since I(x,y, z) = const., the vector (Ix, I", Iz) = A
lies along the positive normal of M. Thus
W=
(A,9u
IAI/I. 9v) du /I. dv = I9u /I. 9v Idu /I. dv.
It is now easy to check that W(9u,9v) = area (9u,9v), hence W(Vl,V2) =
area (Vb V2)'
2) a) Let W = xdy - ydx and j: M <-+ R2 the inclusion of a bounded
region with regular boundary 8M. Show that the area of M is given by
(1/2) JOM rw.
b) Let W = xdy /I. dz - ydx /I. dz + zdx /I. dy and j: MeR3 the inclusion
of a bounded region with regular boundary 8M. Show that the volume
of M is given by (1/3) JOM rw.
c) Generalize the above for R n.
3) Let
xdy /I. dz + ydz /I. dx + zdx /I. dy
(x 2 + y2 + z2)3/2
w=~~~~~~~~~--~
n = 0, if 0 ¢ M, and n = 471", if 0 E M.
where u is the area element of 8 2 and Ll2cp = cpzz + cpyy + cpzz is the
Laplacian of cpo
Hint: Notice that by Euler's relation for homogeneous functions (cf.
Exercise 18, Chapter 1) xcpz+ycpy+zcpz = kcp, and use the divergence
theorem.
b) Let cp = alx4 + a2y4 + a3z4 + 3a4x2y2 + 3aSy2z2 + 3a6x2z2, then
1 52
cp U
471"
= -5 La;.
6
;=1
that is, dw = (rot v, N}u and i*w = (v, t}ds. Now apply Stokes
theorem.
b) Let p E R 3 , ~ a unit vector in R! and P the plane normal to ~,
passing through p and whose orientation together with ~ gives the
orientation of R3. Consider a disk D c P, with center p, and apply
(a) to the surface made up of D and its boundary 8D to obtain
:t
b) H 9 is harmonic in M and
~f (grad9,N) = 0
in oM, where N is the unit normal vector of oM, then 9 = const. in
M.
Hint: Use Green's first identity with f = 9.
c) If 91 and 92 are harmonic in M and
091 092
oN = oN
in oM, then 91 = 92+ const. in M.
d) If 9 is harmonic in M, then
f a9 u = 0
iOM oN
e) The function (Z:a+II:a~%:aP/2 is harmonic in R3 - {o}.
f) (Mean value theorem). Let f be harmonic in the region
Br = {p E R 3 j Ip - Pol 2 $ r2}
whose boundary is the sphere Sr with center in Po. Then
f(Po) = 4:r2l.. fu
Hint: Use Green's second identity in the region D = Br - Bp, P < r,
with f = f and 9 = l/r. Since 9 and f are harmonic,
isf (f~(!)
aN r
-!raNaf ) u = is..
f (f~(!) _! af ) u.
aN r raN
p
{ 7Jii
8R_~=A
Oz
op _ ~ =B
~ of. - C
7f%-F,i-
where P, Q and R are unknown functions in R3.
a) Show that a necessary and sufficient condition for a solution to the
above system to exist is that
8A 8B 8C_ O
8x + 8y + 8z - .
(dei)p(V) = :~:)Wij)p(v)ej.
j
The n2
forms Wij so defined are called the connection forms of R n in the
moving frame {ei}.
Not all of the forms Wij are independent. If we differentiate (ei' ej) = Dij,
we obtain
0= (dei, ej) + (ei' dej) = Wij + Wji,
that is, the connection forms Wij = -Wji are antisymmetric in the indices
i,j.
The crucial point in the method of moving frames is that the forms Wi
and Wij satisfy the structure equations of Elie Cart an.
(2)
Proof. Let al = (1, ... ,0), ... ,an = (0, ... ,1) the canonical basis ofRn, and
let Xi: U - t R be the function that assigns to the point (Xl! ••• , xn) its ith_
coordinate. Then dXi is a differential I-form on U and, since dXj(aj) = Dij,
we conclude that {dXi} is the coframe associated to {ail.
Now write
(4)
]
Wi = L !3ijdxj. (5)
j
1. The Structure Equations of R n 79
To obtain the first structure equation (2), we differentiate (5) and use (6):
that is
L dwikf3kj = L Wik 1\ L wksf3sj,
k k s
dw ie = L Wik 1\ Wke,
k
as we wished. o
Remark 1. If we denote by x: U '-+ R n the inclusion map, to say that the
forms Wi are dual to the frame {ei} is equivalent to saying that dx = EWiei.
Intuitively, the expressions that define Wi and Wij, that is,
describe how the moving frame x, el, ... ,en varies as we move (along a curve
x(t» in U. This was how Elie Cartan introduced the method of moving
frames. The structure equations were then consequences of the "necessary"
relations:
d(dx) = 0,
For instance, the first structure equation can be obtained as follows:
hence
dw j = L Wi" Wij·
i
The main idea of Cart an's method to study the geometry of submanifolds
of RN can be described as follows. Let x: Mn -+ Rn+k be an immersion of a
differentiable manifold Mn into the euclidean space R n+k. It is a consequence
of the inverse function theorem that for p E M there exists a neighborhood
U C M of p such that the restriction xlU C M -+ Rn+k is an embedding.
(See Exercise 4 of Chap. 3).
Let V C Rn+k be a neighborhood of x(P) in R n+k such that V n
M = x(U). Assume that V is such that there exists a moving frame
{el' ... , en, en+l, ... , eq } in V with the property that, when restricted to
x(U), the vectors el, ... , en are tangent to x(U)j such a moving frame is said
to be an adapted frame.
In V we have, associated to the frame {ei}, the coframe forms Wi and the
connection forms Wij which satisfy the structure equations (2) and (3). The
map x: U C M -+ V C Rn+k induce forms X*(Wi), X*(Wij) in U. Since x*
commutes with exterior derivation and exterior products, such forms in U
satisfy again the structure equations (2) and (3). It turns out that the local
metric geometry of U C M is all contained in the structure equations, and
this reBects the "universal character" of R n ,
In the next section we will apply the method of moving frames to a simple
but important case, namely, surfaces in R 3 . For that, we will need a few
preliminary lemmas that we establish now.
Proof. We complete the forms Wi into a basis Wl. •.. , Wn Wr+l, ... , Wn of V*
and we write
Oi = LaijWj + Lbi/w/, 1= r + 1, ... ,n.
j I
Lemma 2. Let U eRn and let Wl, ••• ,Wn be linearly independent differential
l-forms in U. Assume that there exists a set of differential l-forms {Wij},
i,j = 1, ... ,n that satisfy the conditions:
dw j = L Wk 1\ Wkj'
dw j = LWk I\ wkj'
k
Then
LWk 1\ (Wkj - Wkj) = 0,
k
Notice that
and, since the Wi are linearly independent, Bt = -Bji' By using the above
symmetries, we obtain finally that
k . . . . k k
Bji = -B'i = -Blk = Bjk = Bkj = -Bij = -Bji = 0,
that is, Wkj = Wkj. o
82 5. Differential Geometry of Surfaces
2. Surfaces in R3
We now apply the method of moving frames to the special case of surfaces
in R3. Let x: M2 -+ R3 be an immersion of a two-dimensional differentiable
manifold in R3. For each point p E M2, an inner product { , }p is defined in
TpM by the rule:
{VllV2}p = (dXp(Vl),dxp(V2)}x(p),
where the inner product in the right hand side is the canonical inner product
of R3. It is straightforward to check that ( , }p is differentiable and defines a
Riemannian metric in M2 to be called the metric induced by the immersion x.
We will study the local geometry of M2 around a point p E M2. Let
U c M be a neighborhood of p such that the restriction xlU is an embedding.
Let V C R3 be a neighborhood of pis R3 such that V n x(M) = x(U), and
that it is possible to choose in V an adapted moving frame ell e2, e3i this
means that, when restricted to x(U), el and e2 are tangent to x(U) (hence
e3 is normal to x(U».
In V we have, associated to the frame {ei}, the coframe forms Wi and
the connection forms Wij = -Wji, i,j = 1,2,3, which satisfy the structure
equations:
dwl = w21\ W2l + w31\ W3l,
Fig. 5.1
We notice the important fact that if M is oriented, the Gauss map can
be defined globally on M.
Now, since dea = Walel + W32e2, we obtain, for all q E U and all v =
aIel + a2e2 E TQm,
de3(v) = _ (hll
h21 h12)
h22
(al) ,
a2
that is, (-h ij ) is the matrix of the differential of the Gauss map ea: U -+ 8 2
in the basis {e 11 e2} which is the geometric interpretation we were looking
for.
Since the matrix (hij) is symmetric, we conclude immediately that the
differential dea: T M -+ T 8 2 of the Gauss map e3: U -+ 8 2 is a selfadjoint
linear map. From a well known result in Linear Algebra, we know that such a
84 5. Differential Geometry of Surfaces
linear map can be diagonalized with real eigenvalues - AI, - A2 and orthogonal
eigenvectors.
It is usual to define the Gaussian curvature K of M in p by
Theorem 1. (Gauss). K only depends on the induced metric of M2; that is,
if x, x': M2 -+ R3 are two immersions with the same induced metrics, then
K(P) = K'(P), p E M, where K and K' are the Gaussian curvatures of the
immersions x and x', respectively.
hence K = K'. o
Gauss theorem means that the Gaussian curvature, the definition of which
made use of the ambient space R 3, only depends on measurements made
on the surface. This led Gauss, around 1827, to imagine the existence of
geometries that were independent of the ambient space. Because he lacked
adequate tools (in particular, the notion of a differentiable manifold), Gauss
did not develop these ideas which were later (1852) taken up by Riemann.
In general, geometric entities on M that can be computed from WI, W2 and
Wl2 depend only on the induced metric in the sense above described, and we
2. Surfaces in R3 85
The image x(U} is a rotation surface with axis Oz whose generating curve
y = h(s}, z = 9(S} is parametrized by the arc length s (Fig. 5.2).
z = h(s}
= g(s}
Fig. 5.2
e2 = dx (!~)
hGV
86 5. Differential Geometry of Surfaces
and this yields the required expression. In particular, for the sphere of radius
R, we have h(s) = Rcos(s/R), hence K = I/R2.
As we mentioned before, to a given immersion x: M2 --+ R 2 we associate
two quadratic forms at each TpM, p EM, which are defined as follows.
The first quadratic form Ip is merely t h" :J.uadratic form associated to the
bilinear form ( , }p, that is,
I = wi +w~.
where, as usually, we dropped the indication of p.
The second quadratic form is defined in an adapted moving frame by
The expression k(n,ea)(p) is called the normal curvature kn(v) of the surface
in the direction v = a'(O) at the point p. Since IIp(v) = kn(v), we have that
kn(v) is the same for all curves a(s) with the same tangent vector v at p.
Thus, collecting the two interpretations, we conclude that
IIp(v) = -(dea(v), v)p = kn(v).
It is known from Linear Algebra that the maximum and minimum of IIp(v),
as v runs the unit circle 8 1 C Tp8, are the eigenvalues -AI, -A2 of (-de3) and
the corresponding vectors generate the eigenspaces of (-dea). The extremal
normal curvatures (-Ad = kb (-A2) = k2 are called the principal curvatures
at p and the corresponding directions are called the principal directions at p.
The importance of the first and second fundamental forms is that they
determine the local geometry of surfaces in R 3 .
In the same vein, we can work out the whole local geometry of surfaces in
R3 (See, for instance, [de] Chapter 3). We stop here, however, in the hope
that the method is sufficiently clear. It only remains to be explained what is
meant by the statement that I and I I determine entirely the local geometry
of a surface in R 3 . This is the content of the next theorem, and the corollary
following it.
Then there exists a rigid motion p: R3 -t R3 such that the restriction plU =
f·
Proof. Let p E U and f(P) E U' . Let T be the translation of R3 that takes
p to p' = f (P) and let R be the rotation of R 3 that takes ei to e~. Set
p = RoT. We will show that 9 = f 0 p-l: p(U) - t U' is the identity in U' ,
and this implies the statement of the theorem.
Since p is an isometry of R 3 , consider the orthonormal moving frame
ej = dp(ej) in p(U). We will denote with an upper index", the entities
associated to the frame {ej} in p(U). By definition, for all q E p(U) and all
v E Tq(p(U»,
(dej)q(v) = L(Wjj)q(v)(ej)q.
j
9 • Wij
I = (f 0 P-1)* Wij
I = (-l)*f*
P WijI = (-1)*
P Wij = Wij'
-
Because q and v are arbitrary, it follows that ei - e~ 0 9 satisfies the system
of ordinary differential equations
where x: p(U) C R3 and x': U' C R3 are the respective inclusions and the
last equality comes from what we just proved. Since the initial conditions in
p(P) are: (x - x' 0 g)(p(P» = 0, we conclude that x = x' 0 g. Since x and x'
are inclusions, this implies that 9 is the identity, as we wish~d. 0
lor all p E U and all v E TpU. Then, there exists a rigid motion p: R 3 -t R3
such that plU = I.
exists in U a unique form Wl2 = -W2l such that the two first equations hold.
This is indeed the case.
el = fel + ge2
e2 = gel - Je2.
Lemma 4. If {el, e2} and {el, e2} have the same orientation, then
W12 = W12 - T,
W12 = -W12 - T.
W12 = W12 - T,
and this proves the first part of the lemma. The case in which the orientations
are opposite is analogous. 0
Lemma 5. Let p E U c M be a point and let 'Y: I --+ U be a curve such that
'Y(to) = p. Let CPo = angle (el(P), el(P». Then
cp(t) = it to
(f d9 - 9 df ) dt + CPo
dt dt
is a differentiable junction such that:
coscp(t) = f, sencp(t) = g, cp(to) = CPo, dcp = 'Y· T •
where in the last equality we have used that, since f2 + g2 = 1, f f' + gg' = 0.
Therefore, f cOS!{J + 9 sen!(J = const., and since
f(to)cos!{J(to) + g(to)sen!{J(to) = (f2 + g2)(tO) = 1
we conclude (3).
It follows that
(f - COS!{J)2 + (g - sen!{J)2 = f2 + l- 2f cOS!{J - 2g sen!(J + 1 = 0,
hence
cos!{J(t) = f(t),sen!{J(t) = g(t),
and the lemma follows immediately. o
We are now in a position to develop the intrinsic geometry of surfaces.
The first observation is that from (1) and (2) it follows that in a oriented
surface the 2-form
WI /I. W2 = CA:h /I. W2 = u
does not depend on the choice of frames and is therefore globally defined in
M2. The geometric meaning of the form u is obtained as follows. If VI =
aUel + a12e2, V2 = a2lel + a22e2 are linearly independent vectors at a point
p EM, then
u(VI. V2) = det(aij) = area (VI. V2),
where (VI, V2) denotes the parallelogram generated by VI and V2. Because of
that, u is called the area element of M.
The next object of intrinsic geometry is motivated by Gauss theorem.
Then K(P) does not depend on the choice of frames, and it is called the
Gaussian curvature of M at p.
Proof. Let {eb e2} be another moving frame around p. Assume first that
the orientations of the two moving frames are the same. Then
W12 = W12 - r.
Since r = fdg - = 0, hence dw12 = dW12. It follows that
gdf, dr
-KWI /I. W2 = dw12 = dW12 = -K(;h /I. W2 = -KWI /I. W2
hence K = K, as we wished.
If the orientations an opposite, we obtain
3. Intrinsic Geometry of Surfaces 93
Another entity that does not depend on the choice of frames is the (co-
variant) derivative of vectors.
Lemma 6. The covariant derivative does not depend on the choice of frames.
Proof. Let {el' e2} and {e1' e2} be two orthonormal frames around p. Assume
that they have the same orientation. Then
{
Y1 = f'iiI - g'ih
(5)
Y2 = g'iiI + f'jh
where Y(a(t» = E Yi(t)ej = E y;(t)ei, and f, 9 are differentiable functions
with P + 9 2 = 1. By definition,
VxY = (d~l +W21(X)Y2) el + (d~2 +W12(X)Y1) e2
where the functions are taken at t = O. By using (5), and the facts that
W12 = W12 - T and f f' + gg' = 0, we arrive after a long but straightforward
computation that
Thus the form W12 applied to a vector x is the e2-component of the covariant
derivative V xel'
kg = (0*W12)(:s)'
Proof. el is parallel along 0 if and only if Vel el = O. Since (Vel el, el) = 0,
the last statement is equivalent to
0= (Ve1el,e2) = W12(et},
or to 0*W12 = 0, as we wished. o
3. Intrinsic Geometry of Surfaces 95
Proof. Choose two frames {el,e2} and {el,e2} around a(s) as follows: el =
VI IVI and e2 is normal to el in the positive direction; el = a'(s), and e2
is normal to el in the positive direction. As usual, they are first defined
along a small interval of the curve a about a(s), and then extended to a
neighborhood of a(s) in M. Denote by Wl2 and Wl2 the connection forms
associated to {el, e2} and {el, e2}, respectively.
Now cp is the angle from el to el; cp is only defined up to a constant, but
dcp is well defined, and
dcp = a*wl2 - a*wl2'
Since el is a parallel field along a, a·w12 = O. Also, since el = a'(s), we have
that
as we wished. o
The proof of the above proposition also contains the following interpreta-
tion of the Gaussian curvature in terms of parallel transport. Let p E M2, and
D c M be a neighborhood of p homeomorphic to a disk with smooth bound-
ary aD. Let q E aD and Vo E TqM, lVol = 1, and transport V parallelly
around the closed curve aD. When V returns to q, it makes an angle cp with
the initial position Vo. Parametrizing aD as a(s), where s is the arc length
of aD, and using the frames {el(s) = a'(s),e2(s)}, {el(S) = V(S),e2(S)} as
in the above proof, we obtain
K(P) = lim
D-tp
~D'
area
that is, the Gaussian curvature at p measures how different from the identity
is parallel transport along small circles about p.
EXERCISES
Wl = vEdu, W2 = VGdv.
b) The connection form is given by
_ (VE)vd + (v'G)ud
Wl2 - - ,jG u VE v.
K= _.J_
1 {(..;E)V) + (.,fG)u) }.
EG.,fG v
..;E u
4) Let 8 2 = {(x, y, z) E R 3 j x 2 + y2 + Z2 = I}. Prove that there exists no
differentiable nonzero vector field X on 8 2 •
Hint: Assume the existence of such a field X. Let el = XI IXI e consider
the orthonormal oriented frame {el.e2}. Then dwl 2 = - K WI 1\ W2 = -a,
hence
area 8 2 = [ a = - [ dwl2 = - [ W12 = 0,
JS2 JS2 Jas 2 •
which is a contradiction.
5) Consider R2 with the following inner product: If p = (x,y)
U,v E 1'pR2, then
u·v
(u,v)p = (g(p»2'
where u· v is the canonical inner product of R2 and g: R2 --+ R is a
differentiable positive function. Prove that the Gaussian curvature of this
metric is
K = g(gxx + gyy) - (g; + g~).
6) Let M2 C R 3 be a surface with the induced metric. Let p E M2, x E 1'pM2
and Y be a vector field tangent to M2 . Show that
(u, v) around any point such that the first fundamental form I = du 2 +dv 2 •
Clearly if I is as above, K = O. To prove the converse, proceed as follows:
a) Choose a frame {ell e2} around p EM. Since dw I2 = - K WI /I. W2 = 0,
by Poincare's Lemma, there exists a function 0 defined in a neighbor-
hood V of such that dO = W12,
b) Choose another frame {ell e2} by setting ang (ell ex) = O. Show that
the connection form W12 of this frame vanishes identically,
c) Show that WI2 = 0 implies that dW I = dW2 = 0 and use again
Poincare's Lemma to obtain the required local coordinates.
6. The Theorem of Gauss-Bonnet
and the Theorem of Morse
The considerations of the last chapter were strictly local. However, one of the
most interesting features of differential geometry is the connection between
local properties and properties that depend on the entire surface. One of the
most striking of such properties is the so-called Gauss-Bonnet theorem which
we intend to prove in this section.
In his fundamental work (Considerations on curved surfaces, 1827), Gauss
proved the special case of this theorem for geodesic triangles and foresaw its
importance for the development of differential geometry. The theorem for
more general regions is due to O. Bonnet (Jour. Ecole Polytech. 19 (1848),
1-146). With the advent of Topology, it became soon clear that a global for-
mulation of the Gauss-Bonnet theorem would be an important link between
Geometry and Topology. The extension of this result to higher dimensions
became then an important mathematical problem. After some preliminary
work by Allendoerfer and Weil, a satisfactory solution was obtained in 1944
by 8.8. Chern, as an application of the method of moving frames. We will
come back to that in Remark 2 of this section.
Before starting, we want to make the general remark that any differen-
tiable manifold Mn (Hausdorff and with countable basis) can be given a
Riemannian metric. The proof depends on the existence of a partition of
unity. For the compact case (which is the only one we will use), it suffices
to define arbitrarily an inner product (,}O on each coordinate neighborhood
r(UO) of a finite differentiable structure of Mn, and to set
an open disk in the plane. Notice that the number of isolated singular points
is finite, since M is compact.
To each isolated singular point of X, we are going to associate an integer
to be called the index of X at p, as follows. First, choose a Riemannian
metric on M, and consider the moving frame {el,e2}, where el = X/IXI and
e2 is a unit vector field orthogonal to el and in the orientation of M. This
determines differential forms Wl,W2, W12 in V - {pl. Next, we choose another
moving frame {el' e2}, in the same orientation as before, defined throughout
V, thus obtaining forms WI, W2, W12 in V. The difference
W12 - W12 = T
is defined in V - {p}.
Now consider a simple closed curve C that bounds a compact region of V
containing p in its interior; C will be oriented as the boundary of this region.
By Lemma 5 of Chapter 5, the restriction of T to C is the differential of the
angle tp(t} between el and el along C. Thus
11 T = dtp = 27rl.
Proof. Let Ct and C2 be two simple closed curves around p, as in the def-
inition of index. Assume first that Ct and C2 do not intersect and consider
the annular region b. bounded by C1 and C2. Let 11 be the index computed
with C 1 and 12 be the index computed with C2. By Stokes theorem, and the
fact that dT = 0,
11 - 12 = 2. [ T - 2. [ T = 2. [ dT = 0,
211' lCI 211' l C2 211' 1D.
and this proves the Lemma in this case. If C1 and C2 intersect, we choose a
curve Ca that does not intersect both C 1 and C2 • By applying the above, we
conclude that It = Ia =1 2 • 0
1. The Theorem of Gauss-Bonnet 101
~I~
• 0 ill(
~I\' 1=1
1=2
Fig. 6.1
Lemma 2. The definition of I does not depend on the choice of the frame
{el,e2}' More precisely, let Sr = 8B r be the boundary of a disk of mdius
T and center p, and consideT the frame {el' e2} of the definition. Then, the
limit
1
lim -2 i:ih2 = I
r->O 71'
1 - Sr
exists, and j = I.
Proof. Let Sri' Sr2 be concentric circles, T2 < T1, and let 6. be the annular
region bounded by Sri and Sr2' By Stokes theorem,
1 -1Sri
W12
Sr2
W12 =
J6
f ciW12 ~ 0, as T},T2 ~ 0. (1)
lim -1
r-+O 271'
1 -
Sr
W12 = I
1 Sri
W12 - 271'1 = 1Bq
diJ 12 = -1Brl
KWII\ W2. (2)
(3)
Proof. Let (,)0 and (, h be two Riemannian metrics on M. Let, for t E [0,11,
We can now state and prove the Gauss-Bonnet Theorem in the following
form.
1M
K(7 = 271'Lh
k
i=l
where K is the Gaussian curvature of the metric and (7 is its element of area.
Bi a ball with center Pi which is such that it contains no singular point other
than Pi. From Stokes theorem, we have that
[
1M-UBi
Klih /\ W2 = - [
JM-UBi
dW12 = [
JU(lJB;)
W12 = L[
i JlJB;
W12,
• •
where OBi has the orientation induced by Bi (this is the opposite of the
orientation of M - B i , hence the change of sign in the second equality). Now,
take the limit of the above, as the radii of Bi go to zero, and use Lemma 2
to obtain that
as we wished. o
Notice that the right hand side does not depend on the vector field X
and the left hand side does not depend on the metric. Thus we obtain the
striking conclusion that L: Ii is the same for all vector fields with isolated
singularities, and JM K (J is the same for all Riemannian metrics on M.
k
The number L Ii is called the Euler-Poincare characteristic of M and is
i=l
also denoted by X(M). By the above, X(M) is invariant by diffeomorphisms
and 2~ JM K (J does not depend on the metric on M and it is equal to this
invariant.
Since L:i Ii = X(M) does not depend on the chosen field, we conclude that
V - A +F = X(S).
104 6. The Theorem of Gauss-Bonnet and the Theorem of Morse
Fig. 6.2
Example 1. Let us compute X(M2) for the sphere 8 2 and the torus T.
For the sphere, we choose the metric induced by R3 on 8 2 = {p E
R 3 j Ipi = I} for which K == 1 (cf. Example 1 of Chap. 4). Thus Is Ku =
area 8 2 = 411", hence X(8 2 ) = 2. As a consequence, we see that every tangent
vector field on 8 2 has at least one singularity.
For the torus T, we know that it is possible to introduce in T a metric
with K == 0 (cf. Exercise 1 in Chap. 4). Since X(T) does not depend on the
metric, X(T) = O.
8M and denote their indices by II, ... , Ik. Then, for any Riemannian metric
onM,
f
JM
Ka + f
J OM
kgds = 271" Lh
where kg is the geodesic curvature of 8M and ds is the arc element of 8M.
where i: 8M ---t M is the inclusion map and <p is the angle between el and
el along 8M.
f
JM-UBi
KWI AW2 =- fJ M - UBi
dW 12 =! U8Bi
W12 - f
J OM
i·W12,
hence
f K a + f kgds =
JM JOM
271" Lh
as we wished. 0
The number l: Ii is again an invariant by diffeomorphisms of the (com-
pact, oriented) surface M with boundary 8M, and is again called the Euler-
Poincare characteristic X(M) of M. Let us compute some examples:
i l= l
Therefore,
x(D) = "L.Jli = -2
1
11" 8D
kgds -2 -211"r = 1.
11" r
Remark 4. The Gauss-Bonnet Theorem still holds for compact surfaces with
boundaries and corners, that is, those surfaces for which the boundary fails
to be regular at finitely many points which are called comers. Each corner
qj, j = 1, ... , n, gives rise to an external angle aj (which is the positive
angle made by the tangents at the corner) and this must be added to the
total geodesic curvature, so that the theorem now reads:
The proof involves a certain number of technicalities and we will not enter
into that here.
( ~ 8 2 (fog) )
"(f;lfjJ
A= (0,0),
8~(fog) 82U~g)
Ty8x
2. The Theorem of Morse 107
°
Since at a critical point the first derivatives vanish, it is easily checked that
the fact that det A :I does not depend on the parametrization g.
Nondegenerate critical points are the simplest type of critical points. The
behavior of the function fog = h in a neighborhood of such a point can be
easily described using Taylor's expansion:
°
for f.
b) d> in some neighborhood of pj p is then a point of minimum for f.
On the other hand, if det A < 0, there exist exactly two distinct directions for
which the quadratic form vanishesj for all other directions, either d is positive
(a direction of minimum) or negative (a direction of maximum). Such a point
is called a saddle point.
Now we can state the relation we want to prove
Ag = ( ~~)
(P(fog) lj2(fog)
(0,0).
~ or-
It now suffices to observe that both conditions in the statement of the Propo-
sition are equivalent to det(Ag) # 0. 0
It turns out that simple singularities of vector fields are all isolated.
Consider the map <p: U C R2 --+ R2 given by <p(x, y) = (o:(x, y), (3(x, y)).
Since p is a simple singularity,
From Lemma 4 it follows that it makes sense to talk about the index of
a simple singularity. We will now show that they are easily computable.
Proof. Since the index is local and does not depend on the choice of a
metric, we can take M2 = R 2 with the canonical metric, and assume that
°
p = (0,0) = 0. Thus X: R2 --+ R2 is a differentiable map with X(O) = 0,
and the linear part (the differential) of X at is given by
F(p,t) = {
~
t ' if t '"°
°
(p,t) E R2 x I.
dXo(P), if t =
-11
f( Xl, ... ,X n ) - df(txl, ... ,txn) dt
d
= LXi
n
o
i=l
11 0
t
of
a.(txl, ... ,txn)dt;
x,
then, by setting,
we obtain that hi is differentiable, hi(O, ... ,0) = it(O, ... , 0), and
It follows, by setting
i = 1,2,
where hij, i,j = 1,2, are differentiable functions. Since, even for t = 0,
Remark. The above is just another proof of the Lemma used to prove Theo-
rem 3 in Chapter 2.
With all these preliminaries, the proof of Morse Theorem is almost im-
mediate.
2. The Theorem of Morse 111
EXERCISES
Fig. 6.3
a) /q is differentiable
b) P E M2 is a critical point of /q if and only the straight line pq is normal
to M2.
c) The critical point p is degenerate if and only /q(P) = =
l/ki ,i 1,2,
where kl and k2 are the principal curvatures of M2 at p relative to the
normal pq
References
[CHER] Chern, S.S., A simple intrinsic proof of the Gauss-Bonnet formula for
closed Riemannian manifolds, Annals of Math. 45(1944), 747-752.
[dC] do Carmo, M., Differential Geometry of Curves and Surfaces, Prentice-
Hall,1976.
[FIG] Figueiredo, D., A simplified proof of the divergence theorem, American
Math. Monthly, 71(1964), 619-622.
[FISC) Fischer, G., Mathematical Models, Vieweg, Wiesbaden, 1986.
[H1R] Hirsh, M., Differential Topology, Springer-Verlag, Berlin, 1976.
[KELL] Kellog, 0., Foundations of Potential Theory, Dover Publications, New
York,1954.
[LIM 1] Lima, E., Curso de An8.lise, vol. 2, Projeto Euclides, IMPA, Rio de Janeiro,
2.9. ed. 1985 (in Portuguese).
[LIM 2] Lima, E., Orientability of smooth hypersurfaces and the Jordan Brower
separation theorem, Expo Math. 5 (1987), 283-286.
[MAS] Massey, W., Algebraic Topology, an Introduction, Harcourt Brace, 1968.
[MILN] Milnor, J., Morse Theory, Princeton Press, 1963.
[PIC] Picard, E., Traite d'Analyse, Toure I, Gauthier-Villars, Paris, 3eme ed.
1922.
(WAR] Warner, F., Foundations of Diferentiable Manifolds and Lie Groups, 2nd
edition, Springer, 1986.
Index