On Two Classical Theorems of Algebraic Topology 1971
On Two Classical Theorems of Algebraic Topology 1971
On Two Classical Theorems of Algebraic Topology 1971
W. M. Boothby
To cite this article: W. M. Boothby (1971) On Two Classical Theorems of Algebraic Topology, The
American Mathematical Monthly, 78:3, 237-249, DOI: 10.1080/00029890.1971.11992736
Article views: 1
Prof. Boothby received his Michigan Ph.D. under W. Kaplan. Before his present position at
Washington University, he was on the staff of Northwestern University, and he spent a post-
doctoral fellowship of the American Swiss Foundation at the E.T.H., Zurich. Since then he has
held an NSF fellowship at the I.A.S., Princeton, and a sabbatical year at the University of Geneva,
supported both by the NSF and the Amer. Swiss Found. His main research is in differential geom-
etry and Lie groups. Editor.
237
238 W. M. BOOTHBY [March
sisting of all (xl, . 'Xn) with.L: ~ = 1); and the "cylinder" R xsn-t, which
is an n-dimensional su bmanifold of Rn+l = R X Rn. These manifolds are orient-
able, which means that they possess a covering by coordinate neighborhoods
with the property that whenever two of these neighborhoods overlap, the
differentiable functions giving one set of coordinates in terms of the other
have positive J acobian. If a connected manifold is orientable, then there are
two disjoint classes of such coverings; a choice of one of these classes is said
to orient the manifold. Note that the space Rn is automatically oriented by
the fixed coordinate system associated with it. In the sequel manifold will
always mean connected, orientable, and differentiable manifold.
Throughout this paper we shall use the term differentiable to mean C""
(indefinitely differentiable). Thus a real valued function j(x1, · · · , Xn) on an
open subset UCRn is differentiable if it has continuous partial derivatives of
all orders at every point of U; manifolds are assumed to be covered by coordi-
nate neighborhoods for which change of coordinates on overlapping neighbor-
hoods is C""; and mappings from one manifold to another, unless otherwise
stated, will have expressions in local coordinates which are indefinitely dif-
ferentiable. In particular, any real-valued function on an open subset U of a
manifold M is differentiable if its restriction to each local coordinate neighbor-
hood on M is C"" when expressed in terms of the local coordinates, i.e., as a
function on an open subset of Rn(n=dim M).
These definitions should be fairly familiar, but less familiar is the defini-
tion of a differentiable function f on an arbitrary subset A of a manifold M.
In this case we say thatf is differentiable if it can be extended to a differen-
tiable function!* defined on an open subset Usuch that ACUCM: this re-
duces it to the familiar case above. But there is a difficulty; even a trivial case
-A consisting of a single point of Rn-shows that the values of the deriva-
tives may depend on the extension and thus have no meaning for the function
f itself. To avoid this problem we restrict the arbitrariness of A: we shall say
that a connected subset A of a manifold M is a domain if it is the closure of its
interior, i.e., A= V, where V= Int A is an open subset of M. If f is a differen-
tiable function on a domain A, then any two extensions which are differen-
tiable on an open set U containing A must have identical derivatives of all
orders on the open set VC U and thus on V=A. It follows that the derivatives
of all orders are well-defined, continuous functions on A. In particular, the
restrictions of local coordinates of M to a domain A are differentiable func-
tions on A, hence coordinates on A. Therefore the same covering which orients
M orients any domain A of M. If A CM and BCN are domains of manifolds
M and N, then differentiability of a map F:A~B is defined as for manifolds,
i.e., in terms of local coordinates.
Many of the domains we use are even more restricted. A regular domain D
of a manifold M is a domain which is compact, connected, and whose boun-
dary, denoted aD, is an n-1 dimensional submanifold which divides M into
two components. Examples are: (i) the unit interval I= {xERIO;;:ix;;:i1} in
1971] ON TWO CLASSICAL THEOREMS OF ALGEBRAIC TOPOLOGY 239
I
the manifold M= R, (ii) the unit ball Dn = { (xh · · ·, x,.) ER" LX:= 1} in
the manifold M= R", and (iii) M itself if M is compact-in this case oM is
empty. Regular domains are the domains of integration in the treatment of
Stokes' theorem found in [6]; it is important for this theorem that D is ori-
ented by the restrictions to D of the covering by coordinate neighborhoods
which orients M-moreover, this orientation of D induces an orientation on
oD. Thus the fixed orientation of R" by the single system of coordinates which
covers it also orients D" and its boundary, the unit sphere S"- 1 • Not all the
domains which we consider are regular; it will also be necessary for us to use
spaces which consist of the cartesian product IX D of the unit interval and
a regular domain D of a manifold M, for example, the solid cylinder IX D 2•
Spaces of this type are domains of R X M, a manifold; they differ from regular
domains only in that their boundary is not quite a submanifold. Their use
presents no new difficulties.
Since differentiable functions, coordinates, and maps have meaning for a
domain A of a manifold M, we may also consider differential forms and their
exterior derivatives on A just as we do on a manifold. The forms of degree q
are denoted by IV(A) and the totality of all forms of all degrees by 1\(A).
This set forms an associative algebra over the real numbers (products are
denoted with a wedge /\) and each 1\q(A) is a linear subspace. The 0-forms
are just the differentiable functions on A, and 1\q(A) for q>n=dim M con-
tains only the 0-form (0 of the algebra). The exterior derivative d is a linear
map whose iterate d 2 = 0 and which takes each q-form to a q+ 1 form. In par-
ticular, it assigns to each function (0-form) its differential and to each n-form
the only possible n+l form, namely 0. If aE/\r(A) and fJEA•(A), we have
ai\{J = ( -l)r+•[JI\a and d(ai\{J) = dai\{J + (-1)ral\d{:J.
Thus 1\(A) is not commutative, nor is d a homomorphism. However, these
properties do imply that the kernel of the linear map d, i.e., those forms "'
such that dw = 0, is a subalgebra and that the image of d, i.e., those forms 8
such that 8 = d7J for some 7J, is an ideal within it (forms in the kernel are called
closed and those in the image exact). Finally we mention the fact that any dif-
ferentiable map F:A~B of domains induces an algebra homomorphism
F*: 1\ (B)~ 1\ (A) in the opposite direction. It has the properties (i) F* o d
=d o F*, (ii) under composition of maps: (F o G)*= G* o F*, and (iii) the
identity map induces the identity isomorphism.
A form "' on A is completely determined by its expressions in the local
coordinates, which we now discuss briefly. In local coordinates or on an open
subset of R", the differentials dx1, · · · , dx,. of the coordinate functions form
a set of generators for the algebra, and wE 1\ q(U) may be written
with dx; 1\ dx1 = - dx; 1\ dx;.
They also imply that whenever a map F:A~B, A and B domains of manifolds
of dimensions m and n respectively, is expressed in local coordinates, say with
ya=ya(x1, · · ·, x,.), a=1, ···,m, then F*:l\q(B)~IV(A) is computed in
these coordinates by replacing each dy, by its corresponding expression as a
differential of a function of X1, • • • , x,. and regarding the coefficients as com-
posite functions of the x-coordinates through the y's. For example, if q = 1,
then
F*[ .'E b;(y)dy;] = .'E .'E b;(y(x)) ay·
- ' dx;.
i i OXJ
is more precise since it does not involve local coordinates and thus relieves us
of the necessity of checking independence of coordinates.
More generally, consider an arbitrary q-form won IX D. Using local coor-
dinates as described above, its expression on IX U will have the form
(*) w= .L: a, dt 1\ dx 11 1\ · · · 1\ dx 10 _ 1
1 ••• 10 _ 1 + ,L: b~o 1 ••• ~;0 dx~o 1 1\ · · · 1\ dxk.·
COROLLARY (Poincare's Lemma). ror q > 0 each closed q-form on the closed
ball D" CR" is exact. The same statement holds for its interior, the open ball B".
Proof. Let 8 be a q-form on D" with dO= 0. We define H: I X D"-+ D" by
H(t, x1, · · · , x,.) = (tx1, · · · , tx,.); then H(O, x1o • · · , x,.) = (0, · · · , 0) and
H(1, x1, · · · , x,.) = (x1, · · · , x,.). If we let~ =H*O and adopt our earlier no-
tation: (H*O),=Jio for the 1-parameter family of forms defined by 8 on D",
then we see that (H*8) 0 =0 and (H*O)t=O. Since dH*O=H*dO=O, we have
d~=O. By Lemma 1, dd~=O; it follows that 8 is exact as claimed.
REMARK. From the proof it is clear that the theorem holds for any star-
shaped open set or domain of R".
3. A differentiable version of the theorems. In this section we apply the
preceding lemma, Stokes' theorem, and the facts mentioned earlier about
the behavior of differential forms with respect to mappings in order to estab-
lish coo versions of Theorems I and I I. The following lemma is used in both
proofs:
LEMMA 2. Let {} be the (n -l)-form on Sn-l obtained by restricting to 5"- 1
the form
"
ft = :E (-1)i-lx;dx1 1\ · · · 1\ dx1-1 1\ dxi+1 1\ · · · 1\ dx,.
i-1
f {} =
Jsn-1
f aD"
n = JfD" an = n JfD" dx1 1\ . . . 1\ dx,.,
which is clearly not zero. Thus our assumption that{} is exact leads to a con-
tradiction.
The next lemma is the standard one used in the classical proof. In its
continuous version it is of interest in its own right; it says that a solid ball
cannot be retracted onto its boundary.
LEMMA 3. There is no differentiable map F: D"-+S"- 1 such that on the boun-
dary F is the identity, i.e., such that F(x) = x for every xES"- 1 •
1971] ON TWO CLASSICAL THEOREMS OF ALGEBRAIC TOPOLOGY 243
Proof. We suppose that such a map exists and obtain a contradiction. Let
0 be the form on S"- 1 defined in Lemma 2 and let J:S"- 1---+D" be the inclusion
map. Then Fo J:S"- 1---+S"- 1 is just the identity map of S"- 1• Thus (Fo J)*
=J* oF*, implies J*(F*O) =0. Now F*Q is a form on D" and since dF*O
= F*dQ = 0, it is closed. According to the Corollary to Lemma 1 it is exact,
i.e., we have F*O = d8 for some form 8 on D". This means that 0 = J* ( F*O)
= J*dO = dJ*O so that 0 is also exact. But we have proved in Lemma 1 that
this is not the case. It follows that no such map F exists.
From this lemma we now obtain by the usual method ( [3] or [8]) the dif-
ferentiable version of Theorem I.
THEOREM 1'. If G: D"---+D" is a differentiable map, then there is an xED"
suck that G(x) =.z:.
FIG. 1.
Proof. If there is no such point, then x aud G(x) determine a line; let F(x)
be its point of intersection with S"- 1 as we move along the line in the direc-
tion G(x) to x. For xES"- 1 this implies F(x) =x; so that F:D"---+S"- 1 and F
is the identity on S"- 1 • Moreover, treating x, G(x), etc., as vectors, F(x) is
given by [3]
F(x) = x + >.(x)u(x),
where u(x) is the unit vector (x-G(x))/llx-G(x)ll and }.(x) is the scalar de-
fined by }.(::r:) = -::r:· u+ [1-x·x+(x· u)2]t' 2 ; the dot indicates inner product.
It is easily checked from the geometry (Fig. 1) that the term in brackets is
strictly positive so that }.(x) is C"". Thus F: D"---+S"- 1 is a differentiable map
which is the identity on S"- 1 , contradicting Lemma 3. Hence no such G(x)
exists.
We conclude this section by using Lemma 2 to give the differentiable ver-
sion of Theorem I I.
THEOREM 11'. There is no smooth, nowhere vanishing field of vectors tangent
to an even dimensional sphere.
244 W. M. BOOTHBY [March
By Theorem I' we see that G(x) must have a fixed point. However we have the
following inequalities:
I!G(x) - xll = 11 (F(x) - x) - (F(x) - G(x))ll
~ IIF(x) - xll -IIF(x) - G(x)ll
~ 3e- 2e =e.
This is an obvious contradiction, hence no such F can exist.
Proof of Theorem II. We consider the unit sphere S"- 1 with n-1 even and
suppose that u(x) is a continuous field of tangent vectors which is never zero.
Thus, dividing each vector by its length, we may assume in fact that for each
xES"- 1 the tangent vector u(x) has unit length, i.e., we may assume x· u =0,
and 11 ull = 1 =l!xl!. For x = (x1, · · • , x,.) ER" we let r 2 = L x~ = llxl! 2 ; then we
may extend the component functions u1(x), · · · , u,.(x) of u(x), which are de-
fined only on S"-1, to continuous functions ut(x) on D", in fact, all of R" by
setting u*(O) =0 and u*(x) =ru(x/r) for r.eO. Let v(x) = (v1 (x), · · ·, v,.(x)) be
a differentiable E-approximation to u*(x) on D" with e>O small enough so that
w(x) = v(x) - (x · v(x)) x is never zero on S"- 1• The vector w(x) for x ES"- 1 is
just the orthogonal projection of v(x) onto the tangent hyperplane to S"- 1 at x;
it vanishes only if v(x) is parallel to x, i.e., orthogonal to u(x). Clearly it is
possible to choose E so small that this does not occur. Since w(x) is differentiable,
tangent to S"- 1 , and not zero for any xES"- 1, we have a contradiction to
Theorem 11'. Thus no vector field u(x) of the type assumed is possible, and
Theorem 11 is established.
5. Related topics. In this section we shall try to give the general reader some
insight into the reasons for the importance of Theorems I and 11 by mentioning
a few applications, related theorems, and generalizations. To begin with, we
observe that any space X homeomorphic to D" also has the fixed point property
with respect to continuous maps. Indeed, if H:X~D" is a homeomorphism and
F:x~x is continuous, then Ho F o H- 1 : D"~D· has a fixed point x 0 ; it
follows that y 0 =H- 1 (x 0) is fixed under F. In particular, a compact convex sub-
set of R" will have the fixed point property; using the Brouwer theorem (The-
orem I), it is possible--by a very short and straightforward line of argument-to
extend this result to each continuous map of a closed, convex set of a Banach
space into a compact subset of itself (Schauder-Mazur theorem). This general-
246 W. M. BOOTH BY [March
ized fixed point theorem has been used to prove existence theorems for boundary
value problems of differential equations. This will not surprise anyone who
has seen proofs of the Inverse Function Theorem and of the Existence Theorem
for systems of ordinary differential equations which use the Contracting Map-
ping Theorem as in [9] and [10 ]. (This latter theorem is a very simple fixed point
theorem which asserts that any continuous map F:X-+X of a complete metric
space which contracts distances between pairs of points must have a unique fixed
point; although it is not as strong as Theorem I, it is much easier to prove.)
A very nice account of the Schauder theorem is found in Bers lecture notes [11 ] ,
where he gives as an example the following application (C 0 denotes continuous
and C", k > 0, k times continuously differentiable functions):
(5.1) Let L>O be a real number andj(x1, X2, xa) a bounded continuous func-
tion on R 3 ; then there exists a junction x(t) of class C2 which is a solution of the
boundary value problem:
x'(t) = J(t, x, x') and x(O) = 0 = x(L).
To see how this follows from a fixed point theorem one considers the Banach
space C0 of continuous functions y(t) such that y(O) =0 =y(L) and its subspaces
C1 and C2 of functions of class C 1 and C2 respectively, with suitable norms in all
cases. Define H:C1-+C0 by H[x(t)]=J(t, y(t), y'(t)) and G:Co-+C2 by the for-
mula
formula p(n) =2•+8d. The proof that this many vector fields actually exist
depends on purely algebraic results-basically from linear algebra. We shall
give a hint of how this very geometric theorem is related to algebra by showing
that in the case of sa the maximum possible number of independent vector fields
can be defined, namely 3. This occurs only in two other cases: S 1 and SI, as can
be checked rather easily by finding all solutions of the equation p(n) -1 =n-1.
For the case of 5 1, let X= (xi, x2, Xa, X4) be a point of the unit sphere 5 3 in R'.
If we consider it as a vector, then we are required to find three unit vectors
u;(x), i = 1, 2, 3, which are orthogonal to x, hence tangent to S 3 at x, and which
are linearly independent for each x. The following three vectors which are mu-
tually orthogonal clearly satisfy this requirement:
that K is the unit ball; were we to concede the fact that a continuous function/
on an arbitrary compact set KCR" can be extended to a continuous function on
all of R", then the remainder of the proof would establish the lemma in that case
also.
Now, given E>O, we choose ~>0 so xEK= D" (a compact set) and <~ IIYII
implies IJ(y+x) -J(x) I <E. Define g(x) by
g(x) = J R"
J(y + x)tf>&(y)dy1 · · · dy,..
For each fixed x this integral is finite since cf>&(y) = 0 if IIYII
~ ~. i.e., we could
take the region of integration to be the closed ~-ball around x. Moreover,
smce
I I
Thus g(x) -J(x) <EJ cf>&(y)dy1 · · · dy,. =E.
On the other hand, if we change the variable of integration, letting z=y+x,
we have g(x) = JR" f(z )c/>&(z- x)dzt · · · dz,.. From this expression, the rule for
differentiation under the integral sign, and the fact that cp, is C"", it follows that
g is also C"".
References
1. J. F. Adams, Vector fields on spheres, Bull. Amer. Math. Soc., 68 (1962) 39-41.
2. R. H. Bing, The elusive fixed point property, this MONTHLY, 76 (1969) 119-131.
3. J. Milnor, Topology from the Differentiable Viewpoint, University of Virginia Press,
Charlottesville, 1965.
4. M. Hirsch, A proof of the nonretractability of a cell onto its boundary, Proc. Amer. Math.
Soc., 16 (1963) 364-365.
5. H. Flanders, Differential Forms, Academic Press, New York, 1963.
6. W. H. Fleming, Functions of Several Variables, Addison-Wesley, Reading, Mass., 1965.
7. A. Devinatz, Advanced Calculus, Holt, Rinehart & Winston, New York, 1968.
8. Alexandroff-Hopf, Topologie, Springer Verlag, Berlin, 1935.
9. J. Dieudonne, Foundations of Modern Analysis, Academic Press, New York, 1960.
10. Loomis and Sternberg, Advanced Calculus, Addison-Wesley, Reading, Mass., 1968.
11. L. Bers, Topology (Lecture Notes), New York University, 1957.
12. Witold Hurewicz and Henry Wallman, Dimension Theory, Princeton University Press,
1948.
13. David Hilbert and S. Cohn-Vossen, Geometry and the Imagination, Chelsea, New York,
1952.