Previous Up Next
Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.

1 Basics of Metric Spaces

1.1 Metric Spaces

1.1.1 Metric spaces: definition and examples

In Analysis and Calculus the definition of convergence was based on the notion of a distance between points, namely the standard distance between two real numbers is given by

d(x,y)=|xy|.

Similarly, the distance between two points in the plane, given by

d(x,y)=d((x1,x2),(y1,y2))=
(x1y1)2+(x2y2)2
.

A metric space formalises this notion. This will give us the flexibility to talk about distances on function spaces, for example, or introduce other notions of distance on spaces.

Definition 1 (Metric Space) A metric space (X,d) is a set X together with a function d: X × X → ℝ that satisfies the following properties
  1. d(x,y) ≥ 0; and d(x,y)=0  ⇐⇒ x=y (positive definite);
  2. d(x,y)=d(y,x) (symmetric);
  3. d(x,z) ≤ d(x,y)+d(y,z) (triangle inequality).
The function d is called the metric. The word distance will be used interchangeably with the same meaning.
Example 2
  1. X=ℝ. The standard metric is given by d1(x,y)=|xy|. There are many other metrics on , for example
    d(x,y)=|exey|;
    d(x,y)=


    |xy| if  |xy| ≤ 1, 
        1 if  |xy| ≥ 1.
  2. Let X be any set whatsoever, then we can define the discrete metric
    d(x,y) =


          1 if  x ≠ y, 
          0 if  x=y.
  3. X=ℝm. The standard metric is the Euclidean metric: if x=(x1,x2,…,xm) and y=(y1,y2,…,ym) then
      d2(x,y)=
    (x1y1)2+(x2y2)2+…+(xmym)2
    .
    This is linked to the inner-product (scalar product), x ·y=x1 y1+ x2 y2+…+xm ym, since it is just √(xy).(xy). We will study inner products more carefully later, so for the moment we won’t prove the (well-known) fact that it is indeed a metric.

    Other possible metrics include

    d(x,y)=max{|x1y1|,|x2y2|,…,|xmym|}.

    Another metric on ℝm comes from the generalisation of our first example:

      d1(x,y)=|x1y1|+|x2y2|+…+|xmym| .

    These metrics d1, d2, d are all translation-invariant (i.e., d(x+z,y+z)=d(x,y)), and positively homogeneous (i.e., d(kx,ky)=|k|d(x,y)), see Ex. 8 for further discussion.

  4. Take X=C[a,b]. Here are three metrics similar to above ones:
    d2(f,g)=
    b
    a
     | f(x)−g(x) |2   dx
    .

    Again, this is linked to the idea of an inner product, so we will delay proving that it is a metric.

    d1(f,g)=
    b
    a
     |f(x)−g(x)|   dx,

    the area between two graphs

    d(f,g)=max{ |f(x)−g(x)|: a ≤ x ≤ b},

    the maximum vertical separation between two graphs.

Example 3 On C[0,1] take f(x)=x and g(x)=x2 and calculate
d2(f,g)=



1
0
 (xx2)2   dx


1/2



 
=
1/30
,
d1(f,g)=
1
0
 |xx2|   dx = 1/6,  and
d(f,g)=
 
max
x ∈ [0,1]
|xx2|=1/4.
Remark 4 Any subset of a metric space is again a metric space its own right, by restricting the distance function to the subset.
Example 5
  1. The interval [a,b] with d(x,y)=|xy| is a subspace of .
  2. The unit circle {(x1,x2) ∈ ℝ2: x12+x22=1 } with d2(x,y)=√(x1y1)2+(x2y2)2 is a subspace of 2.
  3. The space of polynomials P is a metric space with any of the metrics inherited from C[a,b] above.
Definition 6 A normed space (V,||· ||) is a real vector space V with a map ||·||: V → ℝ (called norm) satisfying
  1. ||v|| ≥ 0, and (||v||=0 ⇔ v=0),
  2. ||λ v|| = |λ| || v|| ,
  3. ||v+w|| ≤ ||v||+ ||w||.
Exercise 7 Prove that V is a metric space with metric d(v,w):=||vw||.
Exercise 8
  1. Write norms ||·||1, ||·||2, ||·|| on m which produces metrics d1, d2, d from Ex. 2.3.
    Hint: see (11) and  (9) below.
  2. Show, that the following are norms on the vector space V=C[a,b]:
         
          || f ||1
    = 
    b
    a
     |f(x)|   dx,
             
    || f ||2
    = 
    b
    a
     |f(x)|2   dx,
             
    || f ||
    = 
     
    sup
    x∈[a,b]
     |f(x)|.
             
    Furthermore, these norms generate the respective metrics d1, d2 and d from Ex. 2(4) as indicated in the previous exercise.
Definition 9 An inner product space(V,⟨·, ·⟩) is a real vector space V with a map ⟨·, ·⟩: V × V → ℝ (called inner product) satisfying
  1. ⟨ λ v,w ⟩ = λ⟨ v,w,
  2. v1 +v2 ,w ⟩ = ⟨ v1,w ⟩+⟨ v2,w,
  3. v,w ⟩ = ⟨ w,v,
  4. v,v ⟩ ≥ 0, and (v,v ⟩ =0 ⇔ v=0).
Exercise 10
  1. Prove that the Cauchy–Schwarz inequality | ⟨ v,w ⟩ |2 ≤ ⟨ v,v ⟩ ⟨ w,w holds.
    Hint: start by considering the expression v + λ w ,v + λ w⟩ ≥ 0 and analyse the discriminant of the quadratic expression for λ.
  2. Then prove that V is a normed space with norm || v||:= ⟨ v,v1/2.
  3. Which of the above norms ||·||1, ||·||2, ||·|| from Ex. 8 can be obtained from an inner product as described in the previous item?

There is a natural name for a class of maps, which preserve metrics:

Definition 11 (Isometry) Let (X, dX) and (Y, dY) be two metric spaces. A map φ: XY is an isometry if
    dY(φ(x1), φ(x2)) =  dX(x1, x2)     for all  x1, x2 ∈ X.
A metric space (X,dX) is isometric to a metric space (Y,dY) if there is an isometry bijection between X and Y.

1.1.2 Open and closed sets

Definition 12 (Open and closed balls) Let (X,d) be a metric space, let xX and let r>0. The open ball centred at x, with radius r, is the set
Br(x)={y ∈ X: d(x,y)<r },
and the closed ball is the set
Br(x)
={y ∈ X: d(x,y) ≤ r }.

Trivial but useful observations are:

Note, that in ℝ with the usual metric the open ball is Br(x)=(xr,x+r), an open interval, and the closed ball is Br(x)=[xr,x+r], a closed interval.

For the d2 metric on ℝ2, the unit ball, B1(0), is disc centred at the origin, excluding the boundary. You may like to think about what you get for other metrics on ℝ2. What are balls in the discrete metric, Ex. 2.2?

Definition 13 (Open sets) A subset U of a metric space (X,d) is said to be open, if for each point xU there is an r>0 such that the open ball Br(x) is contained in U (“room to swing a cat").

Clearly X itself is an open set, that is the whole metric space is open in itself. Also the empty set ∅ is also considered to be open in a trivial way.

Remark 14 Note that the property “be open” of a set depends on the metric space. For example if we consider the set [0,1] it is open in the metric space [0,1] with the standard metric, but not open in the set with standard metric.
Proposition 15 Every “open ball" Br(x) is an open set.
Proof. For if yBr(x), choose δ=rd(x,y). We claim that Bδ(y) ⊂ Br(x).

If zBδ(y), i.e., d(z,y)<δ, then by the triangle inequality

d(z,x) ≤ d(z,y)+d(y,x) <  δ + d(x,y) = r.

So zBr(x). □

Definition 16 (Closed set) A subset F of (X,d) is said to be closed, if its complement XF is open.

Note that closed does not mean “not open". In a metric space the sets ∅ and X are both open and closed. In ℝ we have:

Remark 17 As it can be seen from the definitions the property of a subset F to be open or closed depends from the surrounding space X. For example:
Example 18 If we take the discrete metric,
d(x,y)=


        1 if  x ≠ y, 
        0 if  x=y,
then each point {x}=B1/2(x) so is an open set. Hence every set U is open, since for xU we have B1/2(x) ⊆ U. Hence, by taking complements, every set is also closed.
Theorem 19 In a metric space, every one-point set {x0} is closed.
Proof. We need to show that the set U={xX: xx0} is open, so take a point xU. Now d(x,x0)>0, and the ball Br(x) is contained in U for every 0<r< d(x,x0). □
Theorem 20 Let (Uα)α ∈ A be any collection of open subsets of a metric space (X,d) (not necessarily finite!). Then α ∈ A Uα is open. Let U and V be open subsets of a metric space (X,d). Then UV is open. Hence (by induction) any finite intersection of open subsets is open.
Proof. If x ∈ ∪α ∈ A Uα then there is an α with xUα. Now Uα is open, so Br(x) ⊂ Uα for some r>0. Then Br(x) ⊂ ∪α ∈ A Uα so the union is open.

If now U and V are open and xUV, then ∃ r>0 and s>0 such that Br(x) ⊂ U and B(x,s) ⊂ V, since U and V are open. Then B(x,t) ⊂ UV if t ≤ min(r,s). □

Remark 21 Here we used a common property, which is helpful to remember: the minimum of a finite set of positive numbers is always positive (bigger than 0). However, the infimum of an infinite set of positive numbers can be zero, e.g. inf{1/n: n∈ ℕ}=0. Therefore, a transition from a given infinite set to a suitable finite set will be a reacquiring theme in our course, cf. compact set later in the course.

Thereafter, the collection of open sets is preserved by arbitrary unions and finite intersections.

However, an arbitrary intersection of open sets is not always open; for example (−1/n,1/n) is open for each n=1,2,3,…, but ∩n=1(−1/n,1/n)= {0}, which is not an open set.

For closed sets we swap union and intersection.

Theorem 22 Let (Fα)α ∈ A be any collection of closed subsets of a metric space (X,d) (not necessarily finite!). Then α ∈ A Fα is closed. Let F and G be closed subsets of a metric space (X,d). Then FG is closed. Hence (by induction) any finite union of closed subsets is closed.
Proof.

To prove this we recall de Morgan’s laws. We use the notation Sc for the complement XS of a set SX.

x ∉
 
α
Aα
 ⇐⇒ x ∉Aα for all  α,  so  (Aα)c = Aαc.
x ∉
 
α
Aα
 ⇐⇒ x ∉Aα for some  α,  so  (Aα)c = Aαc.

Write Uα= Fαc =XFα which is open. So ∪α ∈ A Uα is open by Theorem 20. Now, by de Morgan’s laws, (∩α ∈ A Fα)c = ∪α ∈ A Fαc. This is just ∪α ∈ A Uα. Since the complement of ∩α ∈ A Fα is open, it is closed.

Similarly, the complement of FG is FcGc, which is the intersection of two open sets and hence open by Theorem 20. Hence FG is closed. □

Infinite unions of closed sets do not need to be closed. An example is

  
n=1
[
1
n
,∞)=(0,∞),

which is open but not closed in ℝ with standard metric.

Definition 23 (Closure of a set) The closure of S, written S, is the smallest closed set containing S, and is contained in all other closed sets containing S.

The above smallest closed set containing S does exist, because we can define

S
 = {F: F ⊃ S  and  F  closed },

the intersection of all closed sets containing S. There is at least one closed set containing S, namely X itself.

Example 24 In the metric space the closure of S=[0,1) is [0,1]. This is closed, and there is nothing smaller that is closed and contains S.
Exercise 25 Give an example of an open ball Br(x) and the respective closed ball Br(x) with the same centre and radius in a metric space X, such that Br(x) is not the closure of Br(x). Note the slight discontent on our notations, which shall not mislead us in future.
Definition 26 (Dense subset) A subset SX is dense in X if S=X.
Theorem 27 The set of rationals is dense in , with the usual metric.
Proof. Suppose that F is a closed subset of ℝ which contains ℚ: we claim that it F=ℝ.

For U=ℝ ∖ F is open and contains no points of ℚ. But an open set U (unless it is empty) must contain an interval Br(x) for some xU, and hence a rational number within it.

Our only conclusion is that U=∅ and F=ℝ, so that =ℝ. □

Definition 28 (Neighbourhood) We say that V is a neighbourhood (nbh) of x if there is an open set U such that xUV; this means that ∃ δ>0 s.t. Bδ(x) ⊆ V. Thus, a set is open precisely when it is a neighbourhood of each of its points.
Example 29 The half-open interval [0,1) is a neighbourhood of every point in it except for 0.
Theorem 30 For a subset S of a metric space X, we have xS iff VS ≠ ∅ for all nhds V of x (i.e., all neighbourhoods of x meet S).
Proof. If there is a neighbourhood of x that doesn’t meet S, then there is an open subset U with xU and US=∅.

But then XU is a closed set containing S and so SXU, and then xS because xU.

Conversely, if every neighbourhood of x does meet S, then xS, as otherwise XS is as open neighbourhood of x that doesn’t meet S. □

Definition 31 (Interior) The interior of S, intS, is the largest open set contained in S, and can be written as
intS = { U: U ⊂ S   and   U  open }.
the union of all open sets contained in S. There is at least open set within S, namely .

We see that S is open exactly when S=intS, otherwise intS is smaller.

Example 32
  1. In the metric space we have int[0,1)=(0,1); clearly this is open and there is no larger open set contained in [0,1).
  2. intℚ = ∅. For any non-empty open set must contain an interval Br(x) and then it contains an irrational number, so isn’t contained in .
Proposition 33 intS=X ∖ (XS).
Proof. By De Morgan’s laws,
intS={ U: U ⊂ S  and  U  open } 
 =X ∖ {Uc: U ⊂ S  and  U  open }  
 =X ∖ {F: F ⊃ (X ∖ S)  and  F  closed } 
 =
X ∖ (
X∖ S
).
This is because US if and only if Uc= (XU) ⊃ (XS). Also F=Uc is closed precisely when U is open. That is, there is a correspondence between open sets contained in S and closed sets containing its complement. □

1.1.3 Convergence and continuity

Let (xn) be a sequence in a metric space (X,d), i.e., x1,x2,…. (Sometimes we may start counting at x0.)

Definition 34 (Convergence) We say xnx (i.e., xn converges to x) if d(xn,x) → 0 as n → ∞.

In other words: xnx if for any ε>0 there exists N∈ℕ such that for all n>N we have d(x,xn) < ε.

This is the usual notion of convergence if we think of points in ℝd with the Euclidean metric.

Theorem 35 Let (xn) be a sequence in a metric space (X,d). Then the following are equivalent:
  1. xnx;
  2. for every open U with xU, there exists an N>0 such that (n>N) xnU;
  3. for every ε>0 there exists an N>0 such that (n>N) xnBε(x).
Proof. 12 If xnx and xU, then there is a ball Bε(x) ⊂ U, since U is open. But xnx so d(xn,x) < ε for n sufficiently large, i.e., xnU for n sufficiently large.

23 is obvious.

Finally, 31. If the 3 condition works for a given ε>0 and large n the inclusion xnBε(x) implies d(xn,x)<ε. □

Theorem 36 Let S be a subset of the metric space X. Then xS if and only if there is a sequence (xn) of points of S with xnx.
Proof. If xS, then for each n we have B1/n(x) ∩ S ≠ ∅ by Theorem 30. So choose xnB1/n(x) ∩ S. Clearly d(xn,x) → 0, i.e., xnx.

Conversely, if xS, then there is a neighbourhood U of x with US=∅. Now no sequence in S can get into U so it cannot converge to x. □

This can also be phrased as follows, characterising closed set in terms of sequences.

Corollary 37 (Closedness under taking limits) A subset YX of a metric space (X,d) is closed if and only if for every sequence (xn) in Y that is convergent in X its limit is also in Y.

Hence, the closure S is obtained from S by adding all possible limit points of sequences in S.

Example 38
  1. Take (ℝ2,d1), where d1(x,y)=|x1y1|+|x2y2|, where x=(x1,x2) and y=(y1,y2), and consider the sequence (1/n,2n+1/n+1). We guess its limit is (0,2). To see if this is right, look at
    d1





    1
    n
    ,
    2n+1
    n+1



    ,(0,2)


    =


    1
    n



    +


    2n+1
    n+1
    −2


    = 
    1
    n
     + 
    1
    n+1
    → 0
    as n → ∞. So the limit is (0,2).
  2. In C[0,1] let fn(t)=tn and f(t)=0 for 0 ≤ t ≤ 1. Does fnf, (a) in d1, and (b) in d?
    (a)
    d1(fn,f)=
    1
    0
    tn   dt = 
    1
    n+1
     → 0
    as n → ∞. So fnf in d1.
    (b)
    d(fn,f)=max{tn: 0 ≤ t ≤ 1}=1 ¬→0
    as n → ∞. So fn ¬→f in d.
    Note: Say
    gng pointwise on [a,b] as n → ∞ if gn(x) → g(x) for all x ∈ [a,b]. If we define g(x)= {
        0 for  0 ≤ x < 1, 
        1 for   x=1,
    then fng pointwise on [0,1]. But gC[0,1], as it is not continuous at 1.
  3. Take the discrete metric
    d0(x,y)=


            1 if  x ≠ y, 
            0 if  x=y.
    Then xnx  ⇐⇒ d0(xn,x) → 0. But since d0(xn,x)=0 or 1, this happens if and only if d0(xn,x)=0 for n sufficiently large. That is, there is an n0 such that xn=x for all nn0.

    All convergent sequences in this metric are eventually constant. So, for example d0(1/n,0) ¬→0.

A result on convergence in ℝm.

Proposition 39 Take 2 with any of the metrics d1, d2 and d. Then a sequence xn=(an,bn) converges to x=(a,b) if and only if ana and bnb.
Proof. A useful observation is that for any xn and x:
    d1(xn,x) ≥ d2(xn,x) ≥ d(xn,x).
If ana and bnb, then for any ε>0 there are Na and Nb such that for N> Na we have | ana |<ε/2 and for n>Nb | bnb |<ε/2. Thus for any n > N=max(Na, Nb):
    ε > 
ana
+
bnb
 =  d1(xn,x) ≥ d2(xn,x) ≥ d(xn,x) ,
which shows the convergence in all three metrics.

To show the opposite, WLOG assume towards a contradiction that an ¬→a, that is, there exits ε>0 such that for any N there exists n>N such that | ana |>ε. Then:

     
    d1(xn,x) ≥ d2(xn,x) ≥    d(xn,x)= max{
ana
,
bnb
}> 
ana
          

showing the divergence in all three norms.

A similar result holds for ℝm in general.

Now let’s look at continuous functions again.

Theorem 40 If fnf in (C[a,b],d), then fnf in (C[a,b],d1).

Informally speaking, d convergence is stronger than d1 convergence.

Proof. d(fn,f)=max{|fn(x)−f(x)|:   axb} → 0 as n → ∞, so, given ε>0 there is an N so that d(fn,f)<ε for nN. It follows that if nN then
d1(fn,f) = 
b
a
 |fn(x)−f(x)|   dx ≤ 
b
a
 ε   dx = ε(ba),
so d1(fn,f) → 0 as n → ∞. □
Remark 41 It is also true that if d(fn,f) → 0 then fnf point-wise on [a,b]. The converse is false, cf. 38(2).

Now we look at continuous functions between general metric spaces.

Definition 42 (Continuity) Let f: (X,dX) → (Y,dY) be a map between metric spaces. We say that f is continuous at xX if for each ε>0 there is a δε,x>0 such that dY(f(x′),f(x)) < ε for all x′∈ X whenever dX(x′,x) < δε,x.

Another way of saying the same is that for every ε>0 there exists a δ>0 such that

f(Bδ(x)) ⊂ Bε(f(x)).

The map f is continuous, if it is continuous at all points of X.

Theorem 43 (Sequential continuity) For f as above, f is continuous at a if and only if, whenever a sequence xna, then f(xn) → f(a).

In short, f is continuous at a if and only if f permutes with the limit:

f


 
lim
n→ ∞
xn


= 
 
lim
n→ ∞
   f
xn
(7)

for any sequence xna.

Proof. Same proof as in real analysis, more or less. If f is continuous at a and xna, then for each ε>0 we have a δ>0 such that dY(f(x),f(a)) < ε whenever dX(x,a) < δ.

Then there’s an n0 with d(xn,a)<δ for all nn0, and so d(f(xn),f(a))<ε for all nn0. Thus f(xn) → f(x).

Conversely, if f is not continuous at a, then there is an ε for which no δ will do, so we can find xn with d(xn,a)<1/n, but d(f(xn),f(a)) ≥ ε. Then xna but f(xn) ¬→f(a). □

But there is a nicer way to define continuity. For a mapping f: XY and a set UY, let f−1(U) be the set, called pre-image or inverse image

f−1(U)={ x ∈ X: f(x) ∈ U }.

This makes sense even if f−1 is not defined as a function.

Theorem 44 (Continuity and open sets) A function f: XY is continuous if and only if f−1(U) is open in X for every open subset UY. In short: the inverse image of an open set is open.
Proof. Suppose that f is continuous, that UY is open, and that x0f−1(U), so f(x0) ∈ U. Now there is a ball Bε(f(x0)) ⊂ U, since U is open, and then by continuity there is a δ>0 such that dY(f(x),f(x0)) < ε whenever dX(x,x0) < δ. This means that for d(x,x0)<δ, f(x) ∈ U and so xf−1(U). That is, f−1(U) is open.

Conversely, if the inverse image of an open set is open, and x0X, let ε>0 be given. We know that Bε(f(x0)) is open, so f−1(B(f(x0),ε)) is open, and contains x0. So it contains some Bδ(x0) with δ>0.

But now if d(x,x0)<δ, we have xBδ(x0) ⊂ f−1(Bε(f(x0))) so f(x) ∈ Bε(f(x0)) and we have d(f(x),f(x0))<ε. □

Remark 45 Note that for f continuous we do not expect f(U) to be open for all open subsets of X, for example f: ℝ → ℝ, f ≡ 0, then f(ℝ)={0}, not open.
Example 46 Let X=ℝ with the discrete metric, and Y any metric space. Then all functions f: XY are continuous! Indeed, in either way:
Exercise 47 Which functions from a metric space X to the discrete metric space are continuous? Which function from the discrete metric space to are continuous?
Proposition 48 Let X and Y be metric spaces.
  1. A function f : XY is continuous if and only if f−1(F) is closed whenever F is a closed subset of Y.
  2. If f: XY and g: YZ are continuous, then so is the composition gf: XZ defined by (gf)(x) = g(f(x)).
Proof.
  1. We can do this by complements, as if F is closed, then U=Fc is open, and f−1(F)=f−1(U)c (a point is mapped into F if and only if it isn’t mapped into U).

    Then f−1(F) is always closed when F is closed  ⇐⇒  f−1(U) is always open when U is open.

  2. Take UZ open; then (gf)−1(U) = f−1(g−1(U)); for these are the points which map under f into g−1(U) so that they map under gf into U.

    Now g−1(U) is open in Y, as g is continuous, and then f−1(g−1(U)) is open in X since f is continuous.

In many cases we may need a stronger notion.

Definition 49 (Uniform continuity) A function f: (X,dX) → (Y, dY) is called uniformly continuous if for each ε>0 there exists δε>0 such that whenever x,x′∈ X satisfy dX(x,x′)≤δε, we have that dY(f(x),f(x′))≤ε.

Note, that here the same δε shall work for all xX. Thus any uniformly continuous function is continuous at every point. On the other hand the function f(x)=1/x on (0,1) is continuous but not uniformly continuous.

1.2 Useful properties of metric spaces

Metric spaces may or may not have some useful properties which we are discussing in the following subsections: completeness and compactness.

1.2.1 Cauchy sequences and completeness

Recall that if (X,d) is a metric space, then a sequence (xn) of elements of X converges to xX if d(xn,x) → 0, i.e., if given ε>0 there exists N such that d(xn,x)< ε whenever nN. Thus, to show that a sequence is convergent from the definition we need to present its limit x which may not belong to the sequence (xn). It would be convenient to deduce convergence of (xn) just through its own properties without a reference to extraneous x. This is possible for complete metric spaces studied in this subsection.

Often we think of convergent sequences as ones where xn and xm are close together when n and m are large. This is almost, but not quite, the same thing in a general metric space.

Definition 50 (Cauchy Sequence) A sequence (xn) in a metric space (X,d) is a Cauchy sequence if for any ε>0 there is an N such that d(xn,xm)<ε for all n, mN.
Example 51 Take xn=1/n in with the usual metric. Now d(xn,xm)=|1/n−1/m|. Suppose that n and m are both at least as big as N; then d(xn,xm) ≤ 1/N. Hence if ε>0 and we take N>1/ε, we have d(xn,xm)≤ 1/N whenever n and m are both N.

In fact all convergent sequences are Cauchy sequences, by the following result.

Theorem 52 Suppose that (xn) is a convergent sequence in a metric space (X,d), i.e., there is a limit point x such that d(xn,x) → 0. Then (xn) is a Cauchy sequence.
Proof. Take ε>0. Then there is an N such that d(xn,x)<ε/2 whenever nN. Now suppose both nN and mN. Then
d(xn,xm) ≤ d(xn,x)+d(x,xm) = d(xn,x)+d(xm,x) < ε/2+ε/2=ε,
and we are done. □
Proposition 53 Every subsequence of a Cauchy sequence is a Cauchy sequence.
Proof. If (xn) is Cauchy and (xnk) is a subsequence, then given ε>0 there is an N such that d(xn,xm) < ε whenever n, mN. Now there is a K such that nkN whenever kK. So d(xnk,xnl)<ε whenever k, lK. □

Does every Cauchy sequence converge?

Example 54
  1. (X,d)=ℚ, as a subspace of with the usual metric. Take x0=2 and define xn+1=xn/2+1/xn. The sequence continues 3/2, 17/12, 577/408,… and indeed the sequence converges in as xnx where x=x/2+1/x, i.e., x2=2. But this isn’t in .

    Thus (xn) is Cauchy in , since it converges to 2 when we think of it as a sequence in . So it is Cauchy in , but doesn’t converge to a point of .

  2. Easier. Take (X,d)=(0,1). Then (1/n) is a Cauchy sequence in X (since it is Cauchy in , as seen above), and has no limit in X.
In each case there are “points missing from X”.
Definition 55 (Completeness) A metric space (X,d) is complete if every Cauchy sequence in X converges to a limit in X.
Theorem 56 The metric space is complete.
Remark 57 In parts of the literature is simply defined as the completion of . In this case one does not have to prove that is complete, but it is complete by construction. One then has to work a bit to show that it is also a field.

This is a result from the first year. Since its proof depends on the definition of ℝ we will not demonstrate it here.

Example 58
  1. Open intervals in are not complete; closed intervals are complete.
  2. What about C[a,b] with d1, d2 or d?

    Following our consideration in Ex. 38.2, define fn in C[0,2] by

      fn(x)=


    xn for  0 ≤ x ≤ 1, 
          1 for   1 ≤ x ≤ 2.

    [DIAGRAM]

    Then

    d1(fn,fm)=
    2
    0
     |fn(x)−fm(x)|   dx
     =
    1
    0
     |xnxm|   dx
     =
    1
    0
     (xmxn)   dx    if   n ≥ m
     =
    1
    m+1
    1
    n+1
     ≤ 
    1
    m+1
     → 0,

    and hence (fn) is Cauchy in (C[0,2],d1). Does the sequence converge?

    If there is an fC[0,2] with fnf as n → ∞, then 02 |fn(x)−f(x)|   dx → 0, so 01 and 12 both tend to zero. So fnf in (C[0,1],d1), which means that f(x)=0 on [0,1] (from an example we did earlier). Likewise, f=1 on [1,2], which doesn’t give a continuous limit.

  3. Similarly, (C[a,b],d1) is incomplete in general. Also it is incomplete in the d2 metric, as the same example shows (a similar calculation with squares of functions). We will see later that it is complete in the d metric.
Remark 59 Note that 2 is also complete with any of the metrics d1, d2 and d; since a Cauchy/ convergent sequence (vn)=(xn,yn) in 2 is just one in which both (xn) and (yn) are Cauchy/ convergent sequences in (cf. Prop. 39).

Similar arguments show that k is also complete for k=1,2,3,…, and (with the same proof as for Corollary) all closed subsets of k are complete.

If a metric space (X,d) is not complete one can always pass to its abstract completion in the following sense.

Proposition 60 (Abstract completion) Any metric space (X,d) is isometric to a dense subspace of a complete metric space, which is called its abstract completion if (X,d).
Proof.[Sketch of proof] We describe a metric space (X,d) in which X is isometric to a dense subset. Consider the space X′ of Cauchy sequences of X. We define an equivalence relation ∼ on X′ by
(xn) ∼ (yn) ⇔ d(xn,yn) → 0.
The set X is defined to be the set of equivalence classes [(xn)]. It has a well defined metric given by
d([(xn)],[(yn)]):= 
 
lim
n → ∞
d(xn,yn).
One checks easily that this is metric and is well defined (does not depend on the chosen representative xn of [(xn)]). Now there is an injective map XX defined by sending x to the constant sequence (x,x,x,…). This map is an isometry. We can therefore think of (X,d) as a subset of (X,d). This subset is dense because every Cauchy sequence can be approximated by a sequence of constant sequences. So the only difficult bit in this construction is to show that (X,d) is complete. We will sketch the construction of a limit here. It turns out that it verifies completeness on a dense set.
Lemma 61 Suppose that (X,d) is a metric space and let YX be a dense set with the property that every Cauchy sequence in Y has a limit in X. Then (X,d) is complete.
Proof. Let (xn) be a Cauchy sequence in X. Now replace xn with another sequence yn in Y such that d(xn,yn)<1/n. Then, by the triangle inequality, yn is again a Cauchy sequence and converges, by assumption, to some xX. Then also xn converges to x. □

Let us turn to the proof of completeness of X′. Suppose that (xn) is a Cauchy sequence in X. Then, in X′ this sequence has the form ((x1,x1,…),(x2,x2,…),(x3,x3,…),…). This sequence has a limit, namely, (xn) itself. □

Exercise 62 (Extension by continuity) Let (X,d) be a metric space and X1 be a dense subset of X. Let f: X1Y be a uniformly continuous function to a complete metric space (Y,d′). Show that there is a unique function f′: XY which satisfies two properties:
  1. restriction of f to X1 coincides with f, that is f′(x)=f(x) for all xX1;
  2. f is continuous on X.
Furthermore, it can be shown that f is uniformly continuous on X. We will call f the extension of f by continuity and will often keep the same letter f to denote f.

There are many important consequences of Ex. 62, in particular the following.

Corollary 63 All abstract completions of a metric space (X,d) are isometric, in other words, the abstract completions is unique up to isometry.

1.2.2 Compactness

Accordingly to a dictionary: compact—closely and firmly united or packed together. For a metric space a meaning of “closely and firmly united” can be defined in several different forms—through open coverings or convergent subsequences—and we will see that these interpretations are equivalent.

An open cover of a metric space (X,d) is a family of open sets (Uα)α ∈ I such that

 
α ∈ I
Uα=X.

A subcover of a cover is a subset I′ ⊂ I of the index set such that (Uα)α ∈ I is still a cover.

Definition 64 (Compactness) A metric space (X,d) is called compact if every open cover has a finite subcover.

Informally: a space is compact if any infinite open covering is excessive and can be reduced just to a finite one. An example of a compact set is [0,1] and example of non-compact—all reals or the open interval (0,1). An importance of this concept is clarified by Rem. 21.

Definition 65 (Sequential Compactness) A metric space (X,d) is called sequentially compact if every sequence (xn)n ∈ ℕ in X has a convergent subsequence.

The limit of a convergent sequence is called the accumulation point of {xn}. It is instructive to compare the definitions of:

     
    x is the limit of {xn}:    ∀ ε>0   ∃ N   ∀ n>N:   d(x, xn) < ε;         
x is an accumulation point of {xn}:     ∀ ε>0   ∀ N   ∃ n>N:   d(x, xn) < ε.          

Thereafter, x is not an accumulation point of {xn} if for some ε>0 and some N for all subsequent n>N we have d(x, xn)>ε.

Informally: a space is sequentially compact if there is no room to place infinite number of points sufficiently apart from each other to avoid their condensation to a limit. Taking the sequence xn=n shows that the set of all reals is not sequentially compact. On the other hand, we know from previous years that bounded closed set in ℝn every sequence has a convergent subsequence. Therefore, bounded closed sets in ℝn are sequentially compact.

Exercise 66 What are compact sets in a discrete metric space? What are sequentially compact sets in a discrete metric space?
Lemma 67 Let (X,d) be a sequentially compact metric space. Then for every ε >0 there exist finitely many points x1,…,xn such that {Bε(xi)∣ i=1,…,n} is a cover.
Proof. Suppose this were not the case. Then there would exist an ε>0 such that for any finite number of points x1,…,xn the collection of balls Bε(xi) does not cover, i.e.
  
n
i=1
Bε(xi) ≠ X.
Starting with n=1 and then inductively adding points that are in the complement of ∪i=1n Bε(xi) we end up with an infinite sequence of points xi such that d(xi,xk) ≥ ε. This sequence cannot have a Cauchy subsequence (required for convergence) in contradiction with the sequential compactness of X. □
Theorem 68 A metric space (X,d) is compact if and only if it is sequentially compact.
Proof. We show the two directions separately.

Compactness implies sequential compactness: Suppose that X is compact and let (xi)i ∈ ℕ be a sequence. We want to show that it has a convergent subsequence. Suppose (xi) did not have a convergent subsequence. Then no point x is an accumulation point. Therefore, for each xX there exists an ε(x)>0 such that only finitely many i ∈ ℕ for which xiBε(x). Since (Bε(x))xX is an open cover it has a finite subcover, that is a finite number of balls with a finite number of xi in each. This contradicts to the infinite number of elements in the sequence (xi).

Sequential compactness implies compactness: This implication is quite tricky. The proof is again by contradiction. Let us assume our space is sequentially compact and there exists a cover Uα that does not have a finite subcover. By the above lemma there are finitely many points x1,…,xN1 such that B1(xi) is a cover. Each of the balls B1(xi) is covered by Uα as well. Since our cover does not have a finite subcover one of the balls B1(xi) does not have a finite subcover. Denote the relevant point xi by z1.

Again there are finitely many points x1,…,xN2 such that B1/2(xi) is a cover of X. The collection of sets B1(z1) ∩ B1/2(xi), with i=1,…,N2 is also a covering of B1(z1). In the same way as before there is at least one of the xi (which we will again call z2), such that B1(z1) ∩ B1/2(z2) can not be covered by a finite subcover of Uα. Continuing like this we construct a sequence of points zi such that none of the sets

   B1(z1) ⋂ B
 
1
2
(z2) ⋂ … ⋂ B
 
1
N
(zN)

can be covered by a finite subcover of Uα.

By assumption the sequence (zi) has a convergent subsequence. Say z is a limit point of that subsequence. Since Uα is an open cover the point z is contained in one of the Uα and of course that means that an open ball Bε(z) around z is contained in Uα for some ε>0.

Now we show that there exits an N ∈ ℕ such that B1/N(zN) is a subset of Uα (this will be the desired contradiction!). Indeed, choose N large enough so that d(zN,z) + 1/N<ε. Then xB1/N(zN) implies that d(x,z) ≤ d(zN,z) + d(x,zN) < d(zN,z) + 1/N<ε. This means in particular that

   B1(z1) ⋂ B
 
1
2
(z2) ⋂ … ⋂ B
 
1
N
(zN)

is a subset of Uα. Thus, there is a subcover of the set B1(z1) ∩ … ∩ B1/N(zN) consisting of one element Uα. This is a contradiction as we constructed the sequence of balls in such a way that these sets cannot be covered by a finite number of the Uα. □

Definition 69 (Boundedness) A subset AX of a metric space is called bounded if there exists x0X and C>0 such that for all xA we have d(x0,x) ≤ C.
Remark 70 One can easily see, using the triangle inequality, that the reference point x0 can be chosen as any point in X. This means if AX is bounded and x0X, then there exist a C>0 such that d(x0,x) ≤ C for any xA.
Theorem 71 Suppose that AX is a compact subset of a metric space. Then A is closed and bounded.
Proof. First we show A is bounded. Choose any x0X and note that the set Bn(x0) indexed by n ∈ ℕ is an open cover of A. Hence, there exists a finite sub-cover< Bn1(x0),…,BnN(x0). Hence, ABC(x0), where C= max{n1,…,cN}. Hence, A is bounded.

Next assume that (xk) is a sequence in A that converges in X. Since A is compact there exists a subsequence that converges in A. Hence, the limit of xk must also be in A. Therefore, A is closed. □

The converse of this statement is not correct in general. It is however famously correct in ℝm.

Theorem 72 (Heine–Borel) A subset K ⊂ ℝm is compact if and only if it is closed and bounded.
Proof. We just need to combine the above statements. We have already shown that compactness implies closedness and boundedness. If K is closed and bounded we know from Analysis that it is sequentially compact. Therefore it is compact. □

As an illustration of further nice properties of compact spaces we mention the following result:

Exercise 73
  1. Any continuous function on a compact set is bounded.
  2. Any continuous function f: KX from a compact space K to a metric space X is uniformly continuous.
Remark 74 Note that there are two different sorts of properties of metric spaces: Completeness and compactness are of the first sort, closedness is of the second, cf. Rem 17.
site search by freefind advanced

Last modified: November 6, 2024.
Previous Up Next