Anatop
Anatop
This document serves as a set of revision materials for the Cambridge Math-
ematical Tripos Part IB course Analysis and Topology in Michaelmas 2019.
However, despite its primary focus, readers should note that it is NOT a verba-
tim recall of the lectures, since the author might have made further amendments
in the content. Therefore, there should always be provisions for errors and typos
while this material is being used.
Contents
1 Uniform Convergence 1
2 Uniform Continuity 6
3 Metric Space 7
4 Completeness 14
5 Topological Spaces 17
6 Connectedness 24
7 Compactness 27
8 Differentiation 31
1 Uniform Convergence
Definition 1.1. A complex sequence xn is said to converge to a complex number
x if ∀ > 0, ∃N ∈ N such that ∀n > N , |x − xn | < .
1
Note that in this example, despite the fact that all of fn are continuous,
even smooth, the resulting limit f needs not be continuous.
Here is another example:
Example 1.2. Let S = R≥0 and let fn (x) = x2 e−nx , then fn → 0 pointwise,
since
x2 x2 x2 x
0 ≤ |fn (x)| = = 2 2 n3 x 3
≤ = →0
enx 1 + nx + n 2x + 6 ··· nx n
as n → ∞.
There is another form of convergence, called uniform convergence, which is
defined as follows:
Definition 1.3. Let S be a set and let fn : S → C be a sequence of functions.
Let f : S → C be a function. We say fn → f uniformly if ∀ > 0, ∃N ∈ N, ∀n >
N, ∀x ∈ S, |f (x) − fn (x)| < .
Note that the only difference between pointwise and uniform convergence is
that the large integer N does not depend on x if the convergence is uniform.
Note also that uniform convergence implies pointwise convergence, but not the
other way around. Although it does not seem to be such a great difference in
definition, in practice, it makes all the difference in the world.
Proposition 1.1. The sequence in the first example, i.e. fn : [0, 1] → R with
fn (x) = xn , does not converge uniformly.
p We can take for example = 1/2. Then for any n ∈ N, we can take
Proof.
x = n 2/3, so that we have
2 1
|fn (x) − f (x)| = |fn (x)| = >
3 2
So the claimed N does not exist. Therefore the sequence fn does not converge
uniformly.
Proposition 1.2. The sequence in the second example, i.e. fn = R≥0 → R
where fn (x) = x2 e−nx , converges absolutely.
Proof. Note that
x2 x2 x2 2
0 ≤ fn (x) = nx
= n 2 x 2 ≤ 2 2
= 2
e 1 + nx + 2 + · · · n x /2 n
p
Therefore for any > 0, we can take N = d 2/e, so for any x ≥ 0, n > N , we
ahve
2 2 2
|f (x) − fn (x)| = |fn (x)| = fn (x) ≤ < 2 ≤ p =
n2 N ( 2/)2
So fn → 0 uniformly.
In fact, although continuous functions may not converge pointwise to a con-
tinuous function, they do converge uniformly to one.
2
Theorem 1.3. Let S ⊂ C be open. Suppose that fn : S → C is a sequence of
continuous functions. If fn → f uniformly, then f is continuous as well.
Informal sketch. Idea: Transfer the nice property of fn to f . Choose large
enough N such that fn − f is arbitratily small for all n > N . We can always
choose x0 close to x where fn (x) close to f (x). Then just use triangle inequality.
”3- proof”
Proof. ∀ > 0, we can choose large enough N such that sup |fn − f | < /3 We
can choose δ > 0 such that |x − x0 | < δ =⇒ |fn (x) − fn (x0 )| < /3.
|f (x) − f (x0 )| ≤ |f (x) − fn (x)| + |f (x0 ) − fn (x0 )| + |fn (x) − fn (x0 )| < 3 =
3
As desired.
Remark. 1. We can use this theorem to show that xn as in the previous example
does not converge uniformly.
2. It is not true that differentiability is preserved under unform convergence.
Theorem 1.4. Let fn : [a, b] → R be all Riemann integrable. Then if it con-
verges uniformly, its limit is also Riemann integrable. Furthermore,
Z b Z b
lim fn (x) dx = lim fn (x) dx
a n→∞ n→∞ a
Recall that a function is Riemann integerable if and only if the upper and
lower sums of f on the interval can be arbitratily close.
Proof. Firstly f is bounded. Since fn are bounded, we can just choose large
enough n such that |fn − f | < 1 and |fn | < M , then |f | ≤ |f − fn | + |fn | <
+ M < M + 1 so f is bounded.
For > 0 choose N such that sup |fn − f | < /(3(b − a)) for any n > N .
Since fn is integrable, there is some disection D of the interval [a, b] such that
UD (fn ) − LD (fn ) < /3. We have
X
|LD (f ) − LD (fn )| = inf f (x) − inf fn (x) (xi+1 − xi ) < /3
x∈[xi ,xi+1 ] x∈[xi ,xi+1 ]
(xi )∈D
3
Corollary 1.5. For uniform convergence, we can swap infinite sums and in-
tegral. That is, if fn : [a, b] → R is a sequence of integrable functions whose
partial sum converges uniformly to some function f , then f is integrable and
Z b ∞ Z
X b
f (x) dx = fn (x) dx
a n=1 a
Proof. Let
n
X
Fn (x) = fk (x)
k=1
so Fn are integrable and Fn → f uniformly. Then we can just apply the pre-
ceding theorem.
Theorem 1.6. Let fn : [a, b] → R be continuously differentable on [a, b]. As-
sume that the sequence of partial sums of fn0 at every point converges uniformly,
and that there is an c ∈ [a, b] such that
∞
X
fn (c)
n=1
where
∞
X
λ= fn (c)
n=1
4
By FTC, f is differentiable and f 0 (x) = g(x). Since g is continuous, f ∈
C 1 ([a, b]). It remains to show that the series sum of fn (x) converges uniformly
to f (x). Let Fn (x) be the partial sum of the series, then by estimating its
difference with f and the fact that the partial sum of derivatives of fn converges
uniformly (use FTC again), we can show that Fn → fn uniformly. [Write details
later]
Definition 1.4. Let fn be a sequence of scalar function on a set S. We say fn
is uniformly Cauchy on S if ∀ > 0, ∃N ∈ N, ∀x ∈ X, ∀n, m > N ,
|f (x) − fn (x)| ≤ |f (x) − fm (x)| + |fm (x) − fn (x)| < 2/2 <
So fn → f uniformly.
So what we did is to fix the x and the n, then let that m tend to infinity,
then we can use the pointwise convergence to give the result. This is how we
get pass the dependence of N on x in the pointwise convergence result.
Corollary 1.8. Let fn be a sequence of scalar functions on S, let
∞
X
Mn
n=1
be convergent with Mn ≥ 0.
If sup |fn | ≤ Mn for any n, then
∞
X
fn
n=1
converges uniformly.
Proof. Let Fn be the partial sum of fn , Fn is uniformly Cauchy due to the
convergence of the series of Mn . Essentially, ∀ > 0, ∃N ∈ N, ∀n, m > N ,
m
X
Mk <
k=n+1
so
m
X
|fn (x)| <
k=n+1
5
Now we consider the power series
∞
X
an (z − a)n
n=0
(an )∞
0 be a sequence of complex number. Let R be the radius of convergence.
Now on the disk |z − a| < R, we consider
∞
X
f (z) = an (z − a)n
n=0
2 Uniform Continuity
Let U ∈ C and f a scalar function on U . We know what continuity means in
the sense of metric space.
Definition 2.1. We say f is uniformly continuous on U if for any > 0, ∃δ >
0, ∀x, y ∈ U , |x − y| < δ =⇒ |f (x) − f (y)| < δ.
Note that the difference between uniform convergence and our initial form
of convergence is that the value of δ does not depend on the point x.
6
Example 2.1. The standard example that a function is continuous but not
uniformly continuous is that f (x) = x2 on R. To see why, observe that we take
= 1, then choose any δ > 0,
(x + δ/2)2 − x2 = δx + δ 2 /4
We can just choose x > 1/δ and the value would exceed 1.
Theorem 2.1. Let f be a scalar function defined on a closed interval [a, b],
then if f is continuous then it is uniformly continuous.
Proof. (Heine-Borel Theorem gives compactness of the interval which provides
a direct proof.)
Assume that it is not the case, so
3 Metric Space
Definition 3.1. Let M be an arbitrary set, a metric on M is a function d :
M × M → R≥0 such that the following properties hold:
1. ∀x, y ∈ M, d(x, y) = 0 ⇐⇒ x = y.
2. ∀x, y ∈ M, d(x, y) = d(y, x).
3. ∀x, y, z ∈ M, d(x, y) + d(y, z) ≥ d(z, x).
The couple (M, d) is called a metric space.
Example 3.1. 1. Let M be R or C, and d(x, y) = |x − y| is called the usual
metric on those sets.
2. Take M = Rn or Cn , then
p
d((x1 , . . . , xn ), (y1 , . . . , yn )) = |x1 − y1 |2 + . . . + |xn − yn |2
7
5. We can have the `p metric for p ≥ 1 where
p
d((x1 , . . . , xn ), (y1 , . . . , yn )) = p |x1 − y1 |p + . . . + |xn − yn |p
6. Let S be a set and M = `∞ S be the set of bounded scalar function on S.
The metric we can take is the uniform metric d(f, g) = supS |f − g|, which is
well-defined since f, g are bounded.
7. Let M be any set, then define
(
1, if x = y
d(x, y) =
0, otherwise
This is called the discrete metric and the space (M, d) the discrete metric space.
8. Let G be a group that is generated by a symmetric set S. Define d(x, y) be
the least integer n ≥ 0 such that n is the least number of generators to get from
X to y.
This develops to the discipline called geometric group theory.
9. Suppose G is a connected (finite) graph, then we can define the distance
between two vertices x, y to be the length of the shortest path from x to y.
10. Riemannian metric in geometry.
11. Take M to be the integers and we fix a prime p. We can define the p-adic
metric dp (x, y) to be 0 if x = y and kx − ykp = p−n where n is the greatest
power of p in the prime factorisation of |x − y|. It is obvious that it is a metric.
The metric space (Z, dp ) is called the p-adic integers.
12. Let M be the set of all functions from N to PR, that is, the set of all sequences.
∞
So for x = (xn ), y = (yn ), we define d(x, y) = n=0 2−n min{1, |xn − yn |}.
As in before, we can construct new objects from old.
Definition 3.2. Let (M, d) be a metric space and N ⊆ M , then (N, d|N ×N )
is a metric space and is called the metric subspace of (M, d). Sometimes we
denote (N, d|N ×N ) by (N, d).
Example 3.2. C[0, 1] be the set of real continuous functions on the unit interval
having the uniform metric is a subspace of l∞ [0, 1].
Note that there are other metrics on C[0, 1], for example
Z 1
d(f, g) = |f (x) − g(x)| dx
0
2
We also have the L metric
s
Z 1
d(f, g) = |f (x) − g(x)| dx
0
8
We need to generalize to topological spaces to have the notion of a quotient
space in a metric/topological space.
We can introduce (uniform) convergence again in any metric spaces. We work
in a metric space (M, d). 1
Definition 3.4. Given a sequence xn in M and a point x ∈ M , we say xn → x
as n → ∞ if for any > 0, ∃N ∈ N, ∀n > N, d(xn , x) < .
Conversely, if such an x exists for a sequence xn , we say xn is convergent.
Lemma 3.1. Assume xn → x and xn → y, then x = y.
Proof. Assume for the sake of contradiction that it is not the case. We let
= d(x, y), we have a large enough N ∈ N such that n ∈ N =⇒ d(xn , x) <
/2, d(xn , y) < /2, so
A contradiction.
Now we can introduce the concept of limit.
Definition 3.5. Let (xn ) be a convergent sequence which converges to x, we
write
lim xn = x
n→∞
(n)
then we can show that a sequence of sequences (xk ) (where each n gives a
sequence) converges to a sequence (xk ) if and only if for each i, x( n)i → xi .
In fact, if we fix a set S, is there always a metric d on RS such that fn → f
under the metric d if fn → f pointwise on S? The answer is no, as we will need
topological tools for it.
6. Consider C[0, 1] under the uniform metric. Surely the function defined by
fn = xn do not converge, but if we equip C[0, 1] with a different metric, for
example the L1 metric, that is,
Z 1
d(f, g) = |f (x) − g(x)| dx
0
9
space M ⊕p M 0 = (M × M 0 , dp ). Then (xn , yn ) → (x, y) if and only if xn →
x, yn → y.
8. Consider the metric subspace N ⊂ M . If a sequence xn converge to x in N ,
then xn → x in M . The converse is not true since we can take N = M \ {x} if
xn → x.
Definition 3.6. Let (M, d), (M 0 , d0 ) be two metric spaces and f : M → M 0 be
a function. We say f is continuous at a ∈ M if
∀ > 0, ∃δ > 0, ∀b ∈ M, d(a, b) < δ =⇒ d0 (f (a), f (b)) <
If f is continuous at every a ∈ N ⊆ M , then we say f is continuous on N .
Note that if f is continuous on M , it is continuous on any N ⊆ M . The
converse, however, is not true.
Example 3.4. We can take both metric spaces to be R, then consider the
function (
1, if x 6= 0
f (x) =
0, otherwise
Then f is continuous on N = R \ {0} but not on M = R.
Proposition 3.2. f is continuous at a if any only if for any sequence xn → a,
we have f (xn ) → f (a).
Proof. If f is continuous at a, then ∀ > 0, we can find some δ > 0 such
that d(a, b) < δ =⇒ d0 (f (a), f (b)) < . Now, find any xn → a, we can
find N ∈ N, ∀n > N, d(a, xn ) < δ, but with the same N , ∀n > N , we have
d0 (f (a), f (xn )) < by the above. So f (xn ) → f (a).
Conversely, if xn → a =⇒ f (xn ) → f (a) but f is not continuous at a, then we
can find > 0 such that ∀δ > 0, there is some x ∈ M such that d(x, a) < but
d0 (f (x), f (a)) > . We may set δn = 1/n and we can obtain the corresponding
xn . Now xn → a but f (xn ) 6→ f (a). This is a contradiction.
The following two corollaries are then obvious.
Corollary 3.3. Let f and g be continuous scalar functions, then f + g, f × g
and f /g (providing that ∀x, g(x) 6= 0) are all continuous.
Corollary 3.4. If f : M → M 0 , g : M 0 → M 00 are both continuous, then g ◦ f
is continuous.
One can also prove them using − δ, which is not hard either.
Example 3.5. 1. Constant, identity (equipping the same metric) and inclusion
(in the sense of metric subspace) functions are continuous.
2. Real and complex polynomials are continuous.
3. The metric function itself is continuous (in fact Lipschitz) with respect to
the dp metric on M × M .
Definition 3.7. A function (M, d) → (M 0 , d0 ) is Lipschitz continuous if there
is some C ≥ 0 such that
∀x, y ∈ M, d0 (f (x), f (y)) ≤ Cd(x, y)
we sometimes call f to be C-Lipschitz.
10
Proposition 3.5. A Lipschitz function is uniformly continuous.
Proof. Trivial.
Definition 3.9. We fix a metric space (M, d), for x ∈ M and r ≥ 0, the open
ball Dr (x) is the set {y ∈ M : d(x, y) < r}.
So xn → x if and only if ∀ > 0, ∃N ∈ N, n > N =⇒ xn ∈ D (x). And
f : M → M 0 is continuous at a ∈ M if and only if ∀ > 0, ∃δ > 0, ∀x ∈ M, x ∈
Dδ (a) =⇒ f (x) ∈ D (f (a)).
Definition 3.10. On (M, d), for x ∈ M and r ≥ 0, the closed ball Br (x) is the
set {y ∈ M : d(x, y) ≤ r}.
Example 3.7. 1. When M is the real numbers, then an open ball is an open
interval and closed ball is an closed interval.
2. In R2 , B1 (0, 0) is the unit disk with boundary in d2 , and an slanted square
in d1 , and a big square in d∞ .
3. If M is discrete, D1 (x) = {x}, B1 (x) = M .
Note that Bs (x) ⊂ Dr (x) ⊂ Br (x) for any s < r.
Definition 3.11. A subset U ⊂ M with x ∈ U is called a neighbourhood of x
(in M ) if there exists some r > 0 with Dr (x) ⊂ U .
11
Proposition 3.7. In a metric space M , the followings are equivalent:
1. xn → x.
2. For any neighbourhood U of x, there is some N ∈ N, ∀n > N, xn ∈ U .
3. For any open set U containing x, there is some N ∈ N, ∀n > N, xn ∈ U .
Proof. 1 =⇒ 2: ∃r > 0, Dr (x) ⊂ U , so we can choose an N ∈ N, ∀n >
N, d(xn , x) < r =⇒ xn ∈ U .
2 =⇒ 3: Immediate by the preceding lemma.
3 =⇒ 1: Given > 0, take U = D (x), then two statement becomes identical.
12
Definition 3.14. A subset A ⊂ M is closed if whenever xn → x in M for some
sequece (xn ) ∈ A, then x ∈ A.
Example 3.9. If (N, d) is discrete, then every subset of N is both open and
closed.
Definition 3.15. Two metrics on a set are equivalent if they give the same
topology.
Note that it is equivalent to say that the teo metrics have the same conver-
gence sequences since they help identify the closed sets. It also means that they
have the same continuous functions, both to and from, any other spaces. Note
that the two metrics induce the same topology if and only if the identity maps
from both spaces are continuous.
Definition 3.16. A map g : M → M 0 is called a homeomorphism if it is a
bijection and both g and g −1 are continuous.
We say g is an isometry if it is bijective and it is isometric.
We say M and M 0 are homeomorphic if there is a homeomorphism between
them. And M , M 0 be isometric if there is an isometry between them.
Remark. 1. Continuous bijections may not be a homeomorphism. Take g : R →
R where the domain is equipped with discrete metric and the codomain with
the usual metric.
2. A surjective isometric function is an isometry.
Example 3.10. 1. (0, 1), (0, ∞) are homeomorphic. Take x 7→ 1/x.
2. R2 and C are isometric.
13
4 Completeness
Definition 4.1. A metric space is called complete if every Cauchy sequence
converges.
Definition 4.2. A subset A ⊂ M where M is a metric space is called bounded
if there is some r > 0 and z ∈ M such that A ⊂ Br (z).
Lemma 4.1. Convergent =⇒ Cauchy =⇒ Bounded.
Proof. Suppose xn → x, then ∀ > 0, we can find some N ∈ N such that
∀n > N, d(x, xn ) < /2, then ∀n, m > N ,
14
Proof. 1. If N is complete, let (xn ) be a sequence in N such that xn → x in M ,
but this means that (xn ) is Cauchy by Lemma 4.1, therefore it is convergent in
N due to completeness, hence x ∈ N due to uniqueness of limit in metric space.
2. Choose any Cauchy sequence (xn ) in N , we know that (xn ) → x for some
x ∈ M due to completeness of M , but since N is closed, x ∈ N as well, so N is
complete.
Theorem 4.5. Let M be a metric space, then the space of bounded continuous
scalar functions in M , Cb (M ) is complete in the uniform metric.
Proof. Cb (M ) is a metric subspace of `∞ (M ) which is complete. But uniform
limit of continuous functions to a continuous function, so Cb (M ) is closed.
To spell out the proof, fix x ∈ M, > 0, we can choose N such that D(fn , f ) <
/3 where D is the uniform metric. Fix any n ≥ N , then since fn is continuous,
∃δ > 0, d(x, y) < =⇒ |fn (x) − fn (y)| < /3. Hence d(x, y) < δ =⇒
|f (x) − f (y)| ≤ |f (x) − fn (x)| + |f (y) − fn (y)| + |fn (x) − fn (y)| < 3/3 = .
Fix some S 6= ∅, a metric space (N, d0 ). Let `∞ (S, N ) be the space of
bounded functions S → N . Then we can define the uniform metric on `∞ (S, N )
defined by D(f, g) = supx∈S d0 (f (x), g(x)).
Now given a metric space (M, d), let Cb (M, N ) be the set fo bounded continuous
functions M → N , then we have
Theorem 4.6. Let S, M, N be as above, assuming that N is complete, then
`∞ (S, N ) is complete under uniform metric, and since Cb (M, N ) is closed in
`∞ (M, N ) hence complete.
Proof. Analogous to the case where M = R or C.
Example 4.2. 1. For any closed and bounded interval [a, b] ∈ R, then con-
tinuous functions on [a, b] are the continuous and bounded functions on [a, b] is
complete under the uniform metric.
Definition 4.4. A map f : M → M 0 is a contraction mapping if f is L-Lipschitz
with L < 1.
Theorem 4.7 (Contraction Mapping Theorem, aka Banach Fixed Point Theo-
rem). If f is a contraction mapping in a nonempty complete metric space, then
f has an unique fixed point.
Note that it is important for the condition listed to be satisfied.
Example 4.3. 1. If we remove the completeness criterion, f : R \ {0} → R \ {0}
defined by f (x) = x/2, then f is a contraction but do not have fixed point.
2. If we remove L < 1, f : R → R by f (x) = x + 1 is 1-Lipschitz but do not
have any fixed point.
3. f (x) = x + 1/x, [1, ∞)
Proof. Fix x0 ∈ M , then define a sequence xn by xn+1 = f (xn ), so xn =
f n (x0 ). We shall show that this sequence is Cauchy. For n ≥ 2, d(xn , xn−1 ) ≤
Ld(xn−1 , xn−2 ) ≤ Ln−1 d(x1 , x0 ) inductively. For m > n,
d(xm , xn ) ≤ d(xn , xn+1 ) + · · · + d(xm−1 , xm )
≤ (Lm−1 + Lm−2 + · · · + Ln )d(x1 , x0 )
Ln
≤ d(x1 , x0 )
1−L
15
The last term, which only depends on the smaller term n, can be as small as we
want when n is large enough, so the sequence is Cauchy.
Hence there is a limit x of the sequence xn since M , but since f is continuous,
f (xn ) → f (x), but f (xn ) = xn+1 , so by uniqueness of limits, f (x) = x.
Suppose f (x) = x and f (y) = y, then if x 6= y, |x − y| = |f (x) − f (y)| ≤
L|x − y| < |x − y| which is a contradiction.
Note that xn → x exponentially fast, so it can also be applied to numerical
analysis to find an approximated solution of the fixed point.
An application of the contraction mapping theorem is to analyze the existence
and uniqueness of the solution of an initial value problem.
Example 4.4. The IVP f 0 (t) = f (t2 ), f (0) = y0 on C[0, 1/2] is what we are
interested in. Assume that f has a solution, then immediately f is continuously
differentiable. By FTC,
Z t
f (t) = f (0) + f (x2 ) dx
0
Let M = C[0, 1/2], which is nonempty and complete, then consider the mapping
Rt
T : M → M defined by (T g)(t) = y0 + 0 g(x2 ) dx T is trivially well-defined
since x ∈ [0, 1/2] =⇒ x2 ∈ [0, 1/4] ⊂ [0, 1/2] and that g(x2 ) is continuous in x.
Also by FTC, (T g)0 = g, so T g is continuously differentiable hence continuous.
Now f solves the IVP iff f is a fixed point of T . Also we can check that T is a
contraction. Indeed, take g, h ∈ M, then
Z t
|T g(t) − T h(t)| = g(x2 ) − h(x2 ) dx
0
Z t
≤ |g(x2 ) − h(x2 )| dx ≤ tD(g, h) ≤ D(g, h)/2
0
16
By FTC, T g is C 1 ((T g)0 (t) = φ(t, g(t))), in particular continuous, whenever g
is continuous. In addition, T g takes values in BR (y0 ) since
Z t Z t
k(T g)(t) − y0 k = φ(x, g(x)) dx ≤ kφ(x, g(x))k dx ≤ C ≤ R
t0 t0
5 Topological Spaces
Definition 5.1. Consider a set X. A topology τ is a collection of subsets of X
such that the following axioms hold:
1. ∅, X ∈ τ . S
2. ∀i ∈ I, Ui ∈ τ =⇒ i∈I Ui ∈ τ .
3. U, V ∈ τ =⇒ U ∩ V ∈ τ .
A topological space is a pair (X, τ ) where X is a set and τ is a topology on X.
Note that the third axiom can be extended to any finite set of elements of
τ.
Members of τ are called open sets of X.
Example 5.1. For any metric space, we can induce the metric topology by
Proposition 3.9. For example, the Euclidean distance on Rn induce the usual
topology on Rn .
17
Definition 5.2. A topological space X (or the topology of X) is called metriz-
able if it can be induced by some metric on X.
In the case where the topology is metrizable, any other metric that is equiv-
alent to the previous metric gives the same topology.
Example 5.2. The indiscrete topology on a set X is {∅, X}.
Definition 5.3. Give topologies τ1 , τ2 on X, we say τ1 is coarser than τ2 or τ2
is finer than τ1 if τ1 ⊂ τ2 .
18
Note again that in a metric space, this reduced to our previous definition.
The proof of this is trivial.
This again and again coincides with previous definition in metric spaces by
Proposition 3.7.
Example 5.6. In a indiscrete space, any sequence converge to any element.
Theorem 5.4. In a Hausdorff space, limits are unique.
19
Note that A◦ ⊂ A ⊂ Ā, and A◦ = A if and only if A is open, Ā = A if and
only if A is closed.
Proposition 5.6.
A◦ = {x ∈ X : A is a neighbourhood of x}
Ā = {x ∈ X : ∀U ⊂ X such that U is a neighbourhood of x, U ∩ A = ∅}
= A ∪ A0
Proof. Exercise.
Example 5.8. In R, [0, 1) ∪ {2} = [0, 1] ∪ {2}, ([0, 1) ∪ {2})◦ = (0, 1), Q̄ =
R, Q◦ = ∅ = Z◦ , Z̄ = Z.
Remark. Convergent sequences determine the metric topology, since x ∈ Ā ⇐⇒
∃(xn ) ∈ A, xn → x. Again we have the ⇐= direction for all topological spaces
but not necessarily for the =⇒ direction.
Definition 5.10. Let X be a topological space and A ⊂ X. We say A is dense
if Ā = X.
We say X is separable if there is a countable dense set in X.
Definition 5.11. 1. Rn is separable since Q̄n = Rn .
2. (non-example) An uncountable set in the discrete topology is not seperable.
Proof. Trivial.
20
Definition 5.13. A base for a topological space (X, τ ) is a family B ⊂ τ such
that ∀U ∈ τ, ∃C ⊂ B such that
[
U= B
B∈C
In other words, the topology τ consists of the arbitrary unions of some family
of open sets which is a subset of B. So a base determines topology.
Example 5.10. 1. The set of all open intervals is a base of the usual topology
on R. In general, the collection of all open balls in a metric space is a base for
the metric topology on it.
However, what we want to do is not to construct B from τ , but the other
way around.
LemmaS5.8. let X be a set and B ⊂ 2X . Assume that
1. X = B∈B B.
2. ∀B1 , B2 ∈ B, ∀x ∈ B1 ∩ B2 , ∃B ∈ B, x ∈ B ⊂ B1 ∩ B2 .
Then there is an unique topology on X that is generated by the base B.
Proof. We must have the topology
( )
[
τ= B:C ⊂B
B∈C
21
Proof. Trivial.
Example 5.12. Constant, identity and inclusion are always continuous. Hence
the restriction of a continuous map is continuous.
Definition 5.16. Let f : X → Y be a map between topological spaces, then we
say f is a homeomorphism if f is a bijection and both f, f −1 are continuous. We
say X, Y are homeomorphic, or X ∼ = Y , if there is a homeomorphism between
them.
Definition 5.17. f is a open map if for every U open in X, f (U ) is open in
Y . So f is a homeomorphism if and only if f is a continuous open bijection.
Definition 5.18. A property P of topological spaces is called a topological
property if it is preserved under homeomorphisms.
Definition 5.19. Let (X, τ ), (Y, ρ) be topological spaces and let
B = {U × V : U ∈ τ, V ∈ ρ}
f −1 (U × V ) = f −1 (U × Y ) ∩ f −1 (X × V )
−1
= f −1 (πX (U )) ∩ f −1 (πY−1 (V ))
= (πX ◦ f )−1 (U ) ∩ (πY ◦ f )−1 (V )
22
Definition 5.20. Start with a topological space (X, τ ) and let R be an equiv-
alence relation on X. We let X/R be the set of equivalence classes (the “quo-
tient set”). Let q : X → X/R be the quotient map sending x 7→ [x] where
[x] = {y ∈ X : yRx} is the equivalence class containing x. The quotient topol-
ogy on X/R is the family
τR = {V ⊂ X/R : q −1 (V ) ∈ τ }
which is open.
∀U, V ∈ τR , q −1 (U ∩ V ) = q −1 (U ) ∩ q −1 (V )
23
Corollary 5.13. If f (x) = f (y) ⇐⇒ xRy, f is surjective, continuous and
open, then f˜ is a homeomorphism.
6 Connectedness
An interval I in R has the defining property that ∀x, y, z, x < y < z, then
x, z ∈ I =⇒ y ∈ I. We know that a real continuous function maps intervals
to intervals due to the intermediate value theorem. But it may not work if the
(restricted) domain is not an interval.
24
Example 6.1. 1. ∅ and singletons are connected.
2. Any indiscrete topological space is connected.
3. The cofinite topology on an infinite set is connected.
4. The discrete topology is disconnected if it is not a singleton.
Lemma 6.3. A subspace Y ⊂ X is disconnected if and only if there are open
sets U, V ∈ X such that U ∩ Y 6= ∅, V ∩ Y 6= ∅, U ∩ V ∩ Y = ∅, Y ⊂ U ∪ V .
Proof. Trivial.
Proposition 6.4. Let Y be a connected subspace of X, then Ȳ is connected.
Proof. Assume not, then by the preceding lemma, there exists open sets U, V
in X such that U ∩ Ȳ 6= ∅, V ∩ Ȳ 6= ∅, U ∩ V ∩ Ȳ = ∅, Ȳ ⊂ U ∪ V . It follows
that U ∩ V ∩ Y = ∅, Y ⊂ U ∪ V , so we must have, WLOG, U ∩ Y = ∅, then
Y ⊂ X \ U =⇒ Ȳ ⊂ X \ U =⇒ Ȳ ∩ U = ∅, contradiction.
Remark. 1. Alternatively, we can use the third part of Theorem 6.2. 2. In fact,
for any Z with Y ⊂ Z ⊂ Ȳ is connected since the closure of Z is Ȳ .
Alternative proof of Lemma 6.1. Let f : X → Y be continuous, for convenience
we can just assume f is surjective using the same argument as the original proof,
then consider any continuous g : Y → Z, then g ◦ f is continuous hence constant
since f is connected, but f is surjective, so g is constant, then it is done by
Theorem 6.2.
Remark. 1. Connectedness is a topological property.
2. If f : X → Y is continuous and A ⊂ X and A is connected, then f (A) is
connected.
Corollary 6.5. If X is connected and R an equivalence relation on X, then
X/R is connected.
Proof. The quotient map is continuous and surjective.
Example 6.2. let Y = {(x, sin(1/x)) : x > 0} ⊂ R2 is connected since it is
the image of f (x) = (x, sin(1/x)), which is continuous since its components are
connected, over R>0 .
By Proposition 6.4, Ȳ = Y ∪ ({0} × [−1, 1]) is also connected. This is called the
Topologist’s Sine Wave.
Lemma 6.6. Let A be a family of connected subset of a topological space X
such that ∀A, B ∈ A , A ∩ B = ∅, then A∈A A is connected.
S
S
Proof. Suppose f : A∈A A → Z is connected, then f |A is continuous for any
A ∈ A , thus it is constant, say it is nAS
, then ∀A, B ∈ A , then nA = nB since
A ∩ B 6= ∅. Thus f is constant, hence A∈A A is connected.
Proposition 6.7. If X, Y are connected, so is X × Y .
Proof. Observe that ∀x ∈ X, {x}×Y ∼ = Y is connected and ∀y ∈ Y, X ×{y} ∼
=X
is connected as well, so since (x, y) ∈ ({x} × Y ) ∩ (X × {y}) 6= ∅, by the
preceding lemma Ax,y = ({x} × Y ) ∪ (X ×S{y}) is connected. Now obviously
(x, y 0 ) ∈ Ax,y ∩ Ax0 ,y0 6= ∅, so X × Y = x∈X,y∈Y Ax,y is connected by the
preceding lemma.
25
Definition 6.2. Let X be a topological space, we define an equivalence relation
R by xRy if and only if there is a connected U ⊂ X such that x, y ∈ U . One
can check that this is an equivalence relation by Lemma 6.6, and the partition
of X by R is called the connected components of X.
Let Cx be the equivalence class containing x.
Proposition 6.8. Connected components are nonempty and are maximal (wrt
inclusion) connected subset of X, also they are closed.
Proof. Let C be a connected component, so it is the equivalence class of some
x, so C = Cx , so C is nonempty since it contains X. So given y ∈ C, ∃Ay 3 x, y
such that U is connected. Ay ∈ C by definition of the S relation. Now ∀y, z ∈
C, x ∈ Ay ∩ Az 6= ∅, therefore by Lemma6.6, hence C = y∈C Ay is connected.
If C ⊂ D and D is connected, then ∀y ∈ D, x, y ∈ D, thus since D is connected
y ∈ C, so D ⊂ C =⇒ C = D.
Hence since C̄ is connected and contains C, by maximality C = C̄, therefore C
is closed.
Definition 6.3. A topological space X is called path-connected if ∀x, y ∈
X, ∃γ : [0, 1] → X continuous, γ(0) = x, γ(1) = y.
Theorem 6.9. Any path-connected space is connected.
Proof. Suppose not, then X is path-connected but not connected, so there are
open U, V disconnects X. Then fixing x ∈ U, y ∈ V , there exists a continu-
ous γ : [0, 1] → X such that γ(0) = x, γ(1) = y. Thus γ −1 (U ), γ −1 (V ) are
nonempty, open, and partitions [0, 1], thus [0, 1] is disconnected by them, which
is a contradiction.
The converse, however, is not true.
Example 6.3. Take the Topologist’s Sine Wave, X = {(x, sin(1/x)) : x >
0} ∪ ({0} × [−1, 1]). We have already shown it is connected. But it is not path-
connected. Indeed, pick points (0, 0), (1, sin(1)) ∈ X. Assume that γ : [0, 1] →
X is continuous and γ(0) = (0, 0) = x, γ(1) = (1, sin(1)) = y. Let γ1 , γ2 be the
components of γ, which are continuous. For γ1 (t) > 0, then [0, γ1 (t)] ⊂ γ1 ([0, t])
by IVT, so ∃n ∈ N, (2πn)−1 , (2πn + π/2)−1 ∈ (0, γ1 (t)) ⊂ γ1 ([0, t]). So there is
some a, b with γ1 (a) = (2πn)−1 , γ1 (b) = (2πn+π/2)−1 , hence γ2 (a) = 0, γ2 (b) =
1, so we can thus find a sequence 1 > t1 > t2 > · · · > 0 with
(
1, if n is even
γ2 (tn ) =
0, otherwise
f −1 (V ) = (f −1 (V ) ∩ A) ∪ (f −1 (V ) ∩ B) = (f |A )−1 (V ) ∪ (f |B )−1 (V )
26
Corollary 6.11. Let X be a topological space. Define the relation R by xRy if
and only if there is a continuous γ : [0, 1] → X such that γ(0) = x, γ(1) = y.
Then this is an equivalence relation.
Proof. Trivial.
Theorem 6.12. Let U ⊂ Rn be open, then U is connected if and only if U is
path-connected.
Proof. It suffice to show every open connected subset of Rn is path-connected.
WLOG U 6= ∅, fix x0 ∈ U , let V be the path-connected component containing
x0 . We shall show that V, U \ V are both open, so by assumption V = U , thus
the proof will be done.
V open: Since U is open, for any x ∈ U , there is r > 0 such that Dr (x) ∈ U .
But any ball is path connected, so ∀x ∈ V, ∃rx > 0, Drx (x) ∈ V , so V is open.
U \ V open: Fix by the same proof as above, any path-connected components
in V is open, so since U \ V is the union of some of them (the ones except V ),
it is open.
Example 6.4. For n ≥ 2, Rn is not homeomorphic to R. Assume f : Rn → R
is a homeomorphism. Fix x ∈ Rn , and let y = f (x), then f |Rn \{x} is still a
homeomorphism to R \ {y}. But then Rn \ {x} is connected by the preceding
theorem, but R \ {y} is not, contradiction.
7 Compactness
Recall that a continuous, real-valued function on a closed bounded interval is
bounded and attains its bound. The question is, for which topological space X
is it true that every continuous real functions is bounded.
Example 7.1. 1. For finite X, every function X → R is bounded.
2. If Sfor all continuous f : X → R, ∃n ∈ N, ∃A1 , A2 , . . . , An ⊂ X such that
X = i Ai and f is bounded on each Ai , then f is bounded on X.
27
with f (x) = m, so for any x ∈ X, f (x) > m so ∃mx such that f (x) > mx > m.
Let Ux = f −1 ((mx , ∞)) which is open and contains x, and inf Ux f ≥ mx > x.
Note that the family of all Ux is an open cover of X, so there is a finite subcover
{Ux }x∈F , so ∀y ∈ X, f (y) ≥ minx∈F mx > m, contradiction.
Note that for a subspace Y ⊂ X, Y is compact iff whenever U is a family
of open set in X whose union contains Y , there is a finite subset V ⊂ U such
that the union of elements in V contains Y .
28
Theorem 7.6 (Topological Inverse Function Theorem). If f : X → Y is a
continuous bijection and X is compact and Y is Hausdorff, then f is a homeo-
morphism.
29
Definition 7.4. Fix a metric space (M, d). For > 0 and F ⊂ M . We say F
is an -net for M if ∀x ∈ M, ∃y ∈ F, d(x, y) ≤ . That is,
[
M= B (y)
y∈F
We say M is totally bounded if for any > 0, there is a finite -net for M .
Note that any compact space is totally bounded, but the converse is not true
by taking [0, 1), but the only thing missing here is completeness.
Theorem 7.9. The followings are equivalent for a metric space M :
(1) M is compact.
(2) M is sequentially compact.
(3) M is totally bounded and complete.
Proof. 1 =⇒ 2: Let (xn ) beTa sequence in M , so for n ∈ N, let An = {xk : k >
n}. It suffices to show that n∈N Ān is nonempty. Assume not, then
[
M \ Ān = M
n∈N
Each
S M \ Ān is open, so by compactness of M there is some N ∈ N such that
n≤N M \ Ān = M . But {An } is decreasing, so necessarily M \ ĀN = M =⇒
ĀN = ∅, contradiction.
2 =⇒ 3: M is complete since a Cauchy sequence with convergent subsequence
is convergent. To see it is totally bounded, assume it is not, then there is some
> 0 such that every -net is infinite. PickSany x1 ∈ M , and once we have
n
already picked x1 , . . . , xn , we pick xn+1 ∈
/ k=1 B (xk ). This is valid since
otherwise M would have a finite -net. But (xn ) cannot possibly have any
Cauchy subsequence, so it has no converging subsequence.
3 =⇒ 1: Assume M is not compact, so there is an open cover U without any
finite subcover. We say A ⊂ M is “bad” ifSthere is no finite subcover of A in
n
U . So M is bad but ∅ is not. Note if A = i=1 Bi is bad, then there is some i
such that Bi is bad.
Next, we want to show that if A is bad and > 0, then ∃B ⊂ A such that B is
bad and diam B = supx,y∈B d(x, y) < . Indeed, since M is bounded, we have a
finite /2-net F , that is,
[ [
B/2 (x) = M =⇒ (B/2 (x) ∩ A) = A
x∈F x∈F
But this would mean that there is some x ∈ F such that B/2 (x) ∩ A is bad, and
by triangle inequality its diameter is less than . Using this we can construct a
sequence M ⊃ A1 ⊃ A2 ⊃ · · · such that An is bad for any n and diam A < 1/n.
Picking xn ∈ An gives a Cauchy sequence (xn ) which converges to some x ∈ M
by completeness. There is some U ∈ U that contains x, and it necessarily
covers An when n is large enough. Contradiction.
Remark. 1. We have a new proof of Bolzano-Weierstrass now! We also have a
new proof of Theorem 7.7 for metric spaces.
2. Sadly, the equivalence of sequentially compactness and compactness fails in
both directions in general topological spaces.
30
8 Differentiation
Definition 8.1. Fix m, n ∈ N, let L(Rm , Rn ) be the set of linear maps to
Rm to Rn . Note that this space is isomorphic to Rmn , both algebraically and
topologically, as we have the metric
v v
uXm Xn um
uX
∀T ∈ L(Rm , Rn ), kT k = t
u
|Tij |2 = t kT ei k2
i=1 j=1 i=1
(b) We have
v v
um um
uX uX
kT Sk = t 2
kT Sei k ≤ t kT k2 kSei k2 = kT kkSk
i=1 i=1
As desired.
Recall that a function f : R → R is differentiable at a if limh→0 (f (a + h) −
f (a))/h exists. So let (h) = (f (a + h) − f (a))/h − f 0 (a), then f (a + h) =
f (a) + f 0 (a)h + (h)h and → 0 as h → 0. We can think of this as (0) = 0
and is continuous at 0. So we want to use it to define differentiation in higher
dimensions.
Definition 8.2. Given m, n ∈ N and an open set U ⊂ Rm , a function f :
U → Rn and a ∈ U . We say f is differentiable at a if there is a linear map
T : Rm → Rn and a function : {h ∈ Rm : a + h ∈ U } → Rn such that
Since U is open, ∃r > 0, Dr (a) ⊂ U , so Dr (a) ⊂ Dom . Note also that our
condition on is also equivalent to say (h)khk = o(khk) as h → 0.
Next, we observe that T (if it exists) is unique. Indeed, if both T, S satisfies
our condition, then (S(h) − T (h))/khk → 0 as h → 0, so by choosing h = x/n
for n ∈ N we have S = T .
31
Definition 8.3. This unique T is called the derivative of f at a, denoted by
f 0 (a) or Df (a) or Df |a , so
Note that f (a, k) + f (h, b) is linear in (h, k), therefore it remains to check
f (h, k) = o(khk). Indeed,
Xm n
X
kf (h, k)k = f hi ei , kj ej
i=1 j=1
X
≤ |hi ||kj |kf (ei , ej )k
i,j
X
≤ k(h, k)k2 kf (ei , ej )k
i,j
f (A + H) = A2 + AH + HA + H 2 = f (A) + AH + HA + o(kHk)
32
Proposition 8.3 (Chain Rule). Consider open U ∈ Rm , V ∈ Rn and functions
f : U → Rn , g : V → Rm and f (U ) ⊂ V . If f is differentiable at a and
g is differentiable at f (a), then g ◦ f is differentiable at a and (g ◦ f )0 (a) =
g 0 (f (a)) ◦ f 0 (a).
Proof. Let b = f (a) and S = f 0 (a), T = g 0 (b), then
(
f (a + h) = f (a) + S(h) + (h)khk
g(b + k) = g(b) + T (k) + δ(k)kkk
33
Pn 0
Since (h) = j=1 j (h)ej has (0) = 0 and is continuous at 0, hence the
result.
Proposition 8.5. Let f, g : U → Rn where U ⊂ Rm is open and φ : U → R is
differentiable at a ∈ U , then so are f + g and φf : x 7→ φ(x)f (x), and
We can do the same thing for products as well which will provide a proof, but we
shall give a different proof. Let F : U → R × Rn = Rn+1 by f (x) = (φ(x), f (x))
and G : R × Rn → Rn by (a, x) 7→ ax. F is differentiable by Proposition 8.4
and G is differentiable since it is bilinear, therefore φf = G ◦ F is differentiable
and we can obtain the form of the derivative from the chain rule which is the
formula as claimed.
Definition 8.5. Let U ⊂ Rm be open and f : U → Rn . Fix a ∈ U and a
direction (nonzero vector) u ∈ Rm \ {0}. The limit
f (a + tu) − f (a)
lim
t→0 t
if exists, is called the directional derivative of f at a to direction u and is denoted
by Du f (a).
Remark. 1. f (a + tu) = f (a) + tDu f (a) + o(t).
2. Let γ(t) = a + tu, then (f ◦ γ)0 (0) = Du f (0).
In the special case where u = ei , we write Di f (a) to denote Dei f (a) and it
is called the ith partial derivative of f at a.
Proposition 8.6. If f is differentiable at a, then all Du f (a) exists and we have
Du f (a) = f 0 (a)(u), so for h = i hi ei , we have
P
X
f 0 (a)(h) = hi Di f (a)
i
Proof. We have
f (a + h) = f (a) + f 0 (a)(h) + (h)khk
Then
f (a + tu) − f (a) ktk
= f 0 (a)(u) + (tu) → f 0 (a)(u)
t t
As t → 0. The rest follows.
34
Remark. 1. Assume f is differentiable at a, then the matrix of f 0 (a) is exactly
represented by (f 0 (a))ji = Di fj (a) = (∂fj /∂xi )(a). This is called the Jacobian
of f at a, denoted by Jf (a).
2. If all partial derivatives exists, so does Du fj (a), ∀j, and we have Du fj (a) =
πj (Du f (a)). So Du πj = πj Du .
3. The converse of the proposition fails in general.
Theorem 8.7. Suppose f : U → Rn where U ⊂ Rm is open. Assume ∃r >
0, Dr (a) ∈ U and Di f exists in Dr (a) and is continuous at a for all i, then f
is differentiable at a.
Proof. WLOG n = 1 by Proposition 8.4 and the second remark above. We shall
prove the case m = 2. The general case is similar.
Let a = (a1 , a2 ) and consider h = (h1 , h2 ) ∈ Dr (0). Certainly we want the
derivative to equal h1 D1 f (a1 , a2 ) + h2 D2 f (a1 , a2 ), so we will try to prove
f (a1 +h1 , a2 +h2 )−f (a1 +h1 , a2 )−h2 D2 f (a1 , a2 ) = φ(h2 )−φ(0)−h2 D2 f (a1 , a2 )
Note that φ is continuous and is differentiable in (−|h2 |, |h2 |). Indeed we have
φ0 (t) = D2 f (a1 + h2 , a2 + t). By MVT, there is some θ(h1 , h2 ) ∈ (0, 1) such that
φ(h2 ) − φ(0) = φ0 (θh2 )h2 . Hence
kf (b) − f (a)k ≤ M kb − ak
35
Corollary 8.9. Let U ⊂ Rm be open and connected, and f : U → Rn be
differentiable such that f 0 ≡ 0, then f is constant.
kx − yk
≥ kh(x) − h(y)k ≥ kx − yk − kf (x) − f (y)k
2
which leads to the desired inequality.
Step 3: For 0 < s < r/2, Ds (0) ⊂ f (B2s (0)) ⊂ f (Dr (0)). Fix y ∈ Ds (0) and
consider h : B2s (0) → Rn by x 7→ y − f (x) + x. We have h0 (x) = −f 0 (x) + I, so
∀x ∈ B2s (0), kh0 (x)k ≤ 1/2. h is then 1/2-Lipschitz by Theorem 8.8. For any
x ∈ B2s (0), kh(x)k = kh(x) − h(0) + yk ≤ kxk/2 + kyk ≤ 2s, so h(B2s (0)) ⊂
B2s (0). By Theorem 4.7 there is some x ∈ B2r (0) such that h(x) = x, which
means that y = f (x).
Step 4: Fix 0 < s < r/2, then let W = Ds (0) and V = f −1 (Ds (0)) ∩ Dr (0), so
V is open and f (V ) = W by step 3 and f is injective by step 2, so f |V : V → W
is a bijection. The inequality in step 2 also implies that the inverse f −1 = g :
W → V is 1/2-Lipschitz, hence continuous.
If g is differentiable, then we have I = (f ◦ g)0 (y) for any y, hence g 0 (y) =
(f 0 (g(y)))−1 by Chain Rule. We want to show that g has this as derivative.
Indeed, fix b ∈ W and a = g(b), T = f 0 (a), we have f (a + h) = f (a) + T (h) +
(h)khk. Fix δ > 0 such that Dδ (b) ∈ W and k ∈ Dδ (0), by setting h =
h(k) = g(b + k) − g(b) we have k = f (a + h) − f (a) = T (h) + (h)khk, so
h = T −1 (k) − T −1 ((h))khk, thus
36
Definition 8.6. Let U ⊂ Rm be open and f : U → Rn . Suppose a ∈ U , then
f is twice differentiable at a if there is some open V with a ∈ V ⊂ U such that
f is differentiable in V and the derivative f 0 : V → L(Rm , Rn ) is differentiable.
f 00 (a) = (f 0 )0 (a) is called the second derivative of f .
So we have f 00 ∈ L(Rm , L(Rm , Rn )) where we have
So the seond derivative is the bilinear map f 00 (A) = B(H, K) = KAH +KHA+
AKH + HKA + AHK + HAK.
Assume that f has second derivative at a under the usual setup, then
Du Dv f (a) = f 00 (a)(u, v)
37
s(Du f (a + αsu + tv) − Du f (a + αsu)). Consider ψ(y) = Du f (a + αsu + yv), so
φ(s, t) = s(ψ(t) − ψ(0)) = stψ 0 (βt), β = β(s, t) ∈ (0, 1). In other words,
φ(s, t)
= Dv Du f (a + αsu + βtv)
st
= f 00 (a + αsu + βtv)(v, u) → f 00 (a)(v, u)
φ(s, t)
→ f 00 (a)(u, v)
st
So they are equal.
Suppose we have U ⊂ Rm open and f : U → R, we say f has a local
maximum at a ∈ U if ∃r > 0, ∀b ∈ Dr (a), f (b) ≤ f (a). We can similarly define
local minima.
Definition 8.7. We say f has a stationary point at a if f is differentiable at a
and f 0 (a) = 0.
It is immediate that f has stationary points at each local maximum and
minimum.
Theorem 8.12. Let U ⊂ Rm be open and f : U → R be twice differentiable
in U and suppose f 0 (a) = 0 and f 00 is continuous at a. If the symmetric form
f 00 is positive definite at a, then f has a local minimum at a; if it is negative
definite at a, then f has a local maximum at a.
38