Anda di halaman 1dari 24

MathCity.

org
Merging man and maths

Metric Spaces
Available online @ http://www.mathcity.org/msc, Version: 2.0

v Metric Spaces
Let X be a non-empty set and denotes the set of real numbers. A function
d : X X is said to be metric if it satisfies the following axioms " x, y, z X .
[M1] d ( x, y ) 0 i.e. d is finite and non-negative real valued function.
[M2] d ( x, y ) = 0 if and only if x = y.
[M3] d ( x, y ) = d ( y, x)
(Symmetric property)
(Triangular inequality)
[M4] d ( x, z ) d ( x, y ) + d ( y, z )
The pair (X, d ) is then called metric space.
d is also called distance function and d(x, y) is the distance from x to y.
Note: If (X, d) be a metric space then X is called underlying set.
v Examples:
i) Let X be a non-empty set. Then d : X X defined by
if x y
1
d ( x, y ) =
if x = y
0
is a metric on X and is called trivial metric or discrete metric.
ii) Let be the set of real number. Then d : defined by
d ( x, y ) = x - y is a metric on .
The space ( , d ) is called real line and d is called usual metric on .
iii) Let X be a non-empty set and d : X X be a metric on X. Then d : X X
defined by d ( x, y ) = min (1, d ( x, y ) ) is also a metric on X.
Proof:
[M1] Since d is a metric so d ( x, y ) 0
as d ( x, y ) is either 1 or d ( x, y ) so d ( x, y ) 0 .
[M2] If x = y then d ( x, y ) = 0 and then d ( x, y ) which is min (1, d ( x, y ) ) will be
zero.
Conversely, suppose that d ( x, y ) = 0 min (1, d ( x, y ) ) = 0
d ( x, y ) = 0
x = y as d is metric.
[M3] d ( x, y ) = min (1, d ( x, y ) ) = min (1, d ( y, x) ) = d ( y, x)
Q d ( x, y ) = d ( y , x )
[M4] We have d ( x, z ) = min (1, d ( x, z ) )
d ( x, z ) 1 or d ( x, z ) d ( x, z )
We wish to prove d ( x, z ) d ( x, y ) + d ( y, z )
now if d ( x, z ) 1 , d ( x, y ) 1 and d ( y, z ) 1
then d ( x, z ) = 1 , d ( x, y ) = 1 and d ( y, z ) = 1
and d ( x, y ) + d ( y, z ) = 1 + 1 = 2

Metric Spaces

therefore d ( x, z ) d ( x, y ) + d ( y, z )
Now if d ( x, z ) < 1 , d ( x, y ) < 1 and d ( y, z ) < 1
Then d ( x, z ) = d ( x, z ) , d ( x, y ) = d ( x, y ) and d ( y, z ) = d ( y, z )
As d is metric therefore d ( x, z ) d ( x, y ) + d ( y, z )
d ( x, z ) d ( x, y ) + d ( y , z )
Q.E.D
iv) Let d : X X be a metric space. Then d : X X defined by
d ( x, y )
d ( x, y ) =
is also a metric.
1 + d ( x, y )
Proof.
d ( x, y )
= d ( x, y ) 0
[M1] Since d ( x, y ) 0 therefore
1 + d ( x, y )
d ( x, y )
[M2] Let d ( x, y ) = 0
= 0 d ( x, y ) = 0 x = y
1 + d ( x, y )
Now conversely suppose x = y then d ( x, y ) = 0 .
d ( x, y )
0
Then d ( x, y ) =
=
=0
1 + d ( x, y ) 1 + 0
d ( x, y )
d ( y, x)
=
= d ( y, x )
[M3] d ( x, y ) =
1 + d ( x, y ) 1 + d ( y , x )
[M4] Since d is metric therefore d ( x, z ) d ( x, y ) + d ( y, z )
a
b
Now by using inequality a < b
<
.
1+ a 1+ b
d ( x, z )
d ( x, y ) + d ( y , z )
We get

1 + d ( x, z ) 1 + d ( x, y ) + d ( y , z )
d ( x, y )
d ( y, z )
d ( x, z )
+
1 + d ( x, y ) + d ( y , z ) 1 + d ( x, y ) + d ( y , z )
d ( x, y )
d ( y, z )
d ( x, z )
+
1 + d ( x, y ) 1 + d ( y , z )
d ( x, z ) d ( x, y ) + d ( y , z )
Q.E.D
v) The space C[a, b] is a metric space and the metric d is defined by
d ( x, y ) = max x(t ) - y (t )
tJ

where J = [a, b] and x, y are continuous real valued function defined on [a, b].
Proof.
[M1] Since x(t ) - y (t ) 0 therefore d ( x, y ) 0 .
[M2] Let d ( x, y ) = 0 x(t ) - y (t ) = 0 x(t ) = y (t )
Conversely suppose x = y
Then d ( x, y ) = max x(t ) - y (t ) = max x(t ) - x(t ) = 0
tJ

tJ

Available at http://www.MathCity.org

Metric Spaces

[M3] d ( x, y ) = max x(t ) - y (t ) = max y (t ) - x(t ) = d ( y, x)


tJ

tJ

[M4] d ( x, z ) = max x(t ) - z (t ) = max x(t ) - y (t ) + y (t ) - z (t )


tJ

tJ

max x(t ) - y (t ) + max y (t ) - z (t )


tJ

tJ

= d ( x, y ) + d ( y , z )
Q.E.D
vi) d : is a metric, where is the set of real number and d defined by
d ( x, y ) =

x- y

vii) Let x = ( x1 , y1 ) , y = ( x2 , y2 ) we define


d ( x, y ) = ( x1 - x2 )2 + ( y1 - y2 ) 2 is a metric on

and called Euclidean metric on 2 or usual metric on 2 .

viii) d : is not a metric, where is the set of real number and d defined by
d ( x, y ) = ( x - y ) 2

Proof.
[M1] Square is always positive therefore ( x - y ) 2 = d ( x, y ) 0
[M2] Let d ( x, y ) = 0 ( x - y ) 2 = 0 x - y = 0 x = y
Conversely suppose that x = y
then d ( x, y ) = ( x - y )2 = ( x - x)2 = 0
[M3] d ( x, y ) = ( x - y ) 2 = ( y - x) 2 = d ( y, x)
[M4] Suppose that triangular inequality holds in d. then for any x, y, z
d ( x, z ) d ( x - y ) + d ( y , z )
( x - z )2 ( x - y ) 2 + ( y - z ) 2
Since x, y, z therefore consider x = 0, y = 1 and z = 2 .

(0 - 2)2 (0 - 1)2 + (1 - 2) 2
4 1+1
42
which is not true so triangular inequality does not hold and d is not metric.

ix) Let x = ( x1 , x2 ) , y = ( y1 , y2 ) 2 . We define


d ( x, y ) = x1 - y1 + x2 - y2
is a metric on 2 , called Taxi-Cab metric on 2 .
x) Let n be the set of all real n-tuples. For
x = ( x1 , x2 ,..., xn ) and y = ( y1 , y2 ,..., yn ) in n
we define d ( x, y ) = ( x1 - y1 ) 2 + ( x2 - y2 ) 2 + ... + ( xn - yn )2
then d is metric on n , called Euclidean metric on n or usual metric on n .
xi) The space l . As points we take bounded sequence
x = ( x1 , x2 ,...) , also written as x = ( xi ) , of complex numbers such that
Available at http://www.MathCity.org

Metric Spaces

xi Cx " i = 1, 2,3,...
where C x is fixed real number. The metric is defined as
d ( x, y ) = sup xi - yi
where y = ( yi )
i

xii) The space l p , p 1 is a real number, we take as member of l p , all sequence


p

x = (x j ) of complex number such that

x
j =1

< .
1
p

p
The metric is defined by d ( x, y ) = x j - h j
j =1

Where y = (h j ) such that

h
j =1

<

Proof.
1
p

p
[M1] Since x j - h j 0 therefore x j - h j = d ( x, y ) 0 .
j =1

[M2] If x = y then

p



p p
p p
p
d ( x, y ) = x j - h j = x j - x j == 0 = 0
j =1

j =1

j =1

Conversely, if d ( x, y ) = 0
1
p

p
x j - h j = 0 x j - h j = 0 (x j ) = (h j ) x = y
j =1

1
p

1
p

p
p
[M3] d ( x, y ) = x j - h j = h j - x j = d ( y, x )
j =1

j =1

[M4] Let z = (z j ) , such that

z
j =1

p
then d ( x, z ) = x j - z j
j =1

<

1
p

p
= x j -h j +hj - z j
j =1

*
Using Minkowskis Inequality

1
p

1
p

p
p
x j -h j + h j - z j
j =1
j =1

= d ( x, y ) + d ( y , z )

1
p

Available at http://www.MathCity.org

Metric Spaces

Q.E.D
v Pseudometric
Let X be a non-empty set. A function d : X X is called pseudometric
if and only if
i) d ( x, x) = 0 for all x X .
ii) d ( x, y ) = d ( y, x) for all x, y X .
iii) d ( x, z ) d ( x, y ) + d ( y, z ) for all x, y, z X .
OR
A pseudometric satisfies all axioms of a metric except d ( x, y ) = 0
may not imply x = y but x = y implies d ( x, y ) = 0 .
Example
Let x, y 2 and x = ( x1 , x2 ) , y = ( y1 , y2 )
Then d ( x, y ) = x1 - y1 is a pseudometric on 2 .
Let x = (2,3) and y = (2,5)
Then d ( x, y ) = 2 - 2 = 0 but x y

Note: Every metric is a pseudometric, but pseudometric is not metric.


* Minkowskis Inequality
If xi = (x1 , x 2 ,..., x n ) and hi = (h1 ,h2 ,...,hn ) are in n and p > 1 , then

xi + hi
i =1

1
p

xi

i =1

1
p


+ hi
i =1

1
p

v Distance between sets


Let ( X , d ) be a metric space and A, B X . The distance between A and B denoted
by d ( A, B) is defined as d ( A, B) = inf {d (a, b) | a A, b B}
If A = { x} is a singleton subset of X, then d ( A, B) is written as d ( x, B ) and is called
distance of point x from the set B.
v Theorem
Let ( X , d ) be a metric space. Then for any x, y X
d ( x, A) - d ( y, A) d ( x, y )
Proof.
Let z A then d ( x, z ) d ( x, y ) + d ( y, z )
then d ( x, A) = inf d ( x, z ) d ( x, y ) + inf d ( y, z )
zA

zA

= d ( x, y ) + d ( y, A)
d ( x, A) - d ( y, A) d ( x, y ) (i)
Next
Available at http://www.MathCity.org

Metric Spaces

d ( y, A) = inf d ( y, z ) d ( y, x) + inf d ( x, z )
z A

zA

= d ( y, x) + d ( x, A)
- d ( x, A) + d ( y, A) d ( y, x)
- ( d ( x, A) - d ( y, A) ) d ( x, y ) (ii )
Combining equation (i) and (ii)
d ( x, A) - d ( y, A) d ( x, y )
Q.E.D

Q d ( x, y ) = d ( y , x )

v Diameter of a set
Let ( X , d ) be a metric space and A X , we define diameter of A denoted by
d ( A) = sup d (a, b)
a , b A

Note: For an empty set j , following convention are adopted


(i) d (j ) = - , some authors take d (j ) also as 0.
(ii) d ( p,j ) = i.e distance of a point p from empty set is .
(iii) d ( A, j ) = , where A is any non-empty set.
v Bounded Set
Let ( X , d ) be a metric space and A X , we say A is bounded if diameter of A is
finite i.e. d ( A) < .
v Theorem
The union of two bounded set is bounded.
Proof.
Let ( X , d ) be a metric space and A, B X be bounded. We wish to prove A B
is bounded.
Let x, y A B
If x, y A then since A is bounded therefore d ( x, y ) <
and hence d ( A B ) = sup d ( x, y ) < then A B is bounded.
x , y A B

Similarity if x, y B then A B is bounded.


Now if x A and y B then
d ( x, y ) d ( x, a) + d (a, b) + d (b, y )
Since d ( x, a), d (a, b) and d (b, y ) are finite
Therefore d ( x, y ) < i.e A B is bounded.

where a A, b B .
Q.E.D

v Open Ball
Let ( X , d ) be a metric space. An open ball in ( X , d ) is denoted by
B( x0 ; r ) = { x X | d ( x0 , x) < r}
x0 is called centre of the ball and r is called radius of ball and r 0 .

Available at http://www.MathCity.org

Metric Spaces

v Closed Ball
The set B( x0 ; r ) = { x X | d ( x0 , x) r} is called closed ball in ( X , d ) .
v Sphere
The set S ( x0 ; r ) = { x X | d ( x0 , x) = r} is called sphere in ( X , d ) .
v Examples
Consider the set of real numbers with usual metric d = x - y " x, y
then B( xo ; r ) = { x | d ( xo , x) < r}
i.e. B( xo ; r ) = { x : x - xo < r}
i.e. x0 - r < x < x + r = ( x0 - r , x0 + r )
i.e. open ball is the real line with usual metric is an open interval.
And B( xo ; r ) = { x : x - x0 r}
i.e. x0 - r x x0 + r = [ x0 - r , x0 + r ]
i.e. closed ball in a real line is a closed interval.
And S ( xo ; r ) = { x : x - x0 = r} = { x0 - r , x0 + r}
i.e. two point x0 - r and x0 + r only.

v Open Set
Let ( X , d ) be a metric space and set G is called open in X if for every x G , there
exists an open ball B( x ; r ) G .
v Theorem
An open ball in metric space X is open.
Proof.
Let B( x0 ; r ) be an open ball in ( X , d ) .
Let y B ( x0 ; r ) then d ( x0 , y ) = r1 < r
Let r2 < r - r1 , then B( y ; r2 ) B ( x0 ; r )
Hence B( x0 ; r ) is an open set.

Alternative:
Let B( x0 ; r ) be an open ball in ( X , d ) .
Let x B( x0 ; r ) then d ( x0 , x) = r1 < r
Take r2 = r - r1 and consider the open ball B ( x ; r2 )
we show that B( x ; r2 ) B ( x ; r ) .
For this let y B ( x ; r2 ) then d ( x, y ) < r2
and d ( x0 , y ) d ( x0 , x) + d ( x, y )
< r1 + r2 = r
hence y B ( x0 ; r ) so that B( x; r2 ) B ( x0 ; r ) . Thus B( x0 ; r ) is an open.
Q.E.D
Available at http://www.MathCity.org

Metric Spaces

Note: Let ( X , d ) be a metric space then


i) X and j are open sets.
ii) Union of any number of open sets is open.
iii) Intersection of a finite number of open sets is open.
v Limit point of a set
Let ( X , d ) be a metric space and A X , then x X is called a limit point or
accumulation point of A if for every open ball B( x ; r ) with centre x,
B( x ; r ) { A - {x}} j .
i.e. every open ball contain a point of A other than x.
v Closed Set
A subset A of metric space X is closed if it contains every limit point of itself.
The set of all limit points of A is called the derived set of A and denoted by A .
v Theorem
A subset A of a metric space is closed if and only if its complement Ac is open.
Proof.
Suppose A is closed, we prove Ac is open.
Let x Ac then x A .
x is not a limit point of A.
then by definition of a limit point there exists an open ball B( x ; r ) such that
B( x ; r ) A = j .
This implies B( x ; r ) Ac . Since x is an arbitrary point of Ac . So Ac is open.
Conversely, assume that Ac is an open then we prove A is closed.
i.e. A contain all of its limit points.
Let x be an accumulation point of A. and suppose x Ac .
then there exists an open ball B( x ; r ) Ac B ( x ; r ) A = j .
This shows that x is not a limit point of A. this is a contradiction to our assumption.
Hence x A . Accordingly A is closed.
The proof is complete.
v Theorem
A closed ball is a closed set.
Proof.
c

Let B( x ; r ) be a closed ball. We prove B ( x ; r ) = C (say) is an open ball.


Let y C then d ( x, y ) > r .
Let r1 = d ( x, y ) then r1 > r . And take r2 = r1 - r
r
r
Consider the open ball B y ; 2 we prove B y ; 2 C .
2
2
Available at http://www.MathCity.org

Metric Spaces

r
r
For this let z B y ; 2 then d ( z, y ) < 2
2
2
By the triangular inequality
d ( x, y ) d ( x, z ) + d ( z , y )
d ( x, y ) d ( z , x ) + d ( z , y )
Q d ( y, z ) = d ( z , y )
d ( z , x ) d ( x, y ) - d ( z , y )
2r - r
2r - r + r
r
r +r
d ( z, x) > r1 - 2 = 1 2 = 1 1
= 1
Q r2 = r1 - r
2
2
2
2
r+r
d ( z, x) >
=r
Q r1 - r = r2 > 0 \ r1 > r
2
z B ( x ; r ) This shows that z C
r
B y; 2 C
2
Hence C is an open set and consequently B( x ; r ) is closed.

Q.E.D

v Theorem
Let ( X , d ) be a metric space and A X . If x X is a limit point of A. then every
open ball B( x ; r ) with centre x contain an infinite numbers of point of A.
Proof.
Suppose B( x ; r ) contain only a finite number of points of A.
Let a1 , a2 ,..., an be those points.
and let d ( x, ai ) = ri where i = 1,2,..., n .
also consider r = min(r1 , r2 ,..., rn )
Then the open ball B( x ; r ) contain no point of A other than x. then x is not limit point of
A. This is a contradiction therefore B( x; r ) must contain infinite numbers of point of A.
v Closure of a Set
Let ( X , d ) be a metric space and M X . Then closure of M is denoted by
M = M M where M is the set of all limit points of M. It is the smallest closed
superset of M.
v Dense Set
Let (X, d) be a metric space the a set M X is called dense in X if M = X .
v Countable Set
A set A is countable if it is finite or there exists a function f : A which is oneone and onto, where is the set of natural numbers.
e.g. , and are countable sets . The set of real numbers, the set of irrational
numbers and any interval are not countable sets.
v Separable Space
A space X is said to be separable if it contains a countable dense subsets.
Available at http://www.MathCity.org

10

Metric Spaces

e.g. the real line is separable since it contain the set of rational numbers, which is
dense is .
v Theorem
Let (X, d) be a metric space, A X is dense if and only if A has non-empty
intersection with any open subset of X.
Proof.
Assume that A is dense in X. then A = X .
Suppose there is an open set G X such that A G = j .
Then if x G then A ( G - {x}) = j
which show that x is not a limit point of A.
This implies x A but x X A X
This is a contradiction.
Consequently A G j for any open G X .
Conversely suppose that A G j for any open G X .
We prove A = X , for this let x X .
If x A then x A A = A then X = A .
If x A then let {Gi } be the family of all the open subset of X such that x Gi for every i.
Then by hypothesis A Gi j for any i. i.e Gi contain point of A other then x.
This implies that x is an accumulation point of A. i.e. x A
Accordingly x A A = A and X = A .
The proof is complete.
v Neighbourhood of a Point
Let (X, d) be a metric space and x0 X and a subset N X is called a
neighbourhood of x0 if there exists an open ball B( x0 ; e ) with centre x0 such that
B( x0 ; e ) N .
Shortly neighbourhood is written as nhood .
v Interior Point
Let (X, d) be a metric space and A X , a point x0 X is called an interior point of
A if there is an open ball B( x0 ; r ) with centre x0 such that B( x0 ; r ) A .
The set of all interior points of A is called interior of A and is denoted by int(A) or Ao .
It is the largest open set contain in A. i.e. Ao A .
v Continuity
A function f : ( X , d ) (Y , d ) is called continuous at a point x0 X if for any
e > 0 there is a d > 0 such that d ( f ( x), f ( x0 ) ) < e for all x satisfying d ( x, x0 ) < d .
Alternative:
f : X Y is continuous at x0 X if for any e > 0 , there is a d > 0 such that
x B( x0 ;d )
f ( x) B ( f ( x0 ); e ) .
Available at http://www.MathCity.org

11

Metric Spaces

v Theorem
f :( X , d ) (Y , d ) is continuous at x0 X if and only if f -1 (G ) is open is X.
wherever G is open in Y.
Note : Before proving this theorem note that if f : X Y , f -1 : Y X and A X ,
B Y then f -1 f ( A) A and f f -1 ( B ) B
Proof.
Assume that f : X Y is continuous and G Y is open. We will prove f -1 (G ) is
open in X.
Let x f -1 (G ) f ( x) f f -1 (G ) G
When G is open, there is an open ball B ( f ( x); e ) G .
Since f : X Y is continuous, therefore for e > 0 there is a d > 0 such that

y B ( x ;d ) f ( y ) B ( f ( x); e ) G then y f -1 f (G ) f -1 (G )
Since y is an arbitrary point of B( x ;d ) f -1 (G ) . Also x was arbitrary, this show that
f -1 (G ) is open in X.
Conversely, for any G Y we prove f : X Y is continuous.
For this let x X and e > 0 be given. Now f ( x) Y and let B ( f ( x); e ) be an open ball

in Y. then by hypothesis f -1 ( B ( f ( x); e ) ) is open in X and x f -1 ( B ( f ( x); e ) )


As y B ( x ;d ) f -1 ( B ( f ( x); e ) )

f ( y ) f f -1 ( B ( f ( x); e ) ) B ( f ( x); e ) i.e. f ( y ) B ( f ( x); e )


Consequently f : X Y is continuous.
The proof is complete.
v Convergence of Sequence:
Let ( xn ) = ( x1 , x2 ,...) be a sequence in a metric space ( X , d ) , we say ( xn ) converges
to x X if lim d ( xn , x) = 0 .
n

We write lim xn = x or simply xn x as n .


n

Alternatively, we say xn x if for every e > 0 there is an n0 , such that


" n > n0 , d ( xn , x) < e .
v Theorem
If ( xn ) is converges then limit of ( xn ) is unique.
Proof.
Suppose xn a and xn b ,
Then 0 d (a, b) d (a, xn ) + d ( xn , b) 0 + 0 as n
d ( a, b) = 0 a = b
Hence the limit is unique.
8
Alternative
Suppose that a sequence ( xn ) converges to two distinct limits a and b.
and d (a, b) = r > 0
Since xn a , given any e > 0 , there is a natural number n1 depending on e
Available at http://www.MathCity.org

12

Metric Spaces

such that

e
whenever n > n1
2
Also xn b , given any e > 0 , there is a natural number n2 depending on e
such that
e
d ( xn , b) <
whenever n > n2
2
Take n0 = max(n1 , n2 ) then
e
e
d ( xn , a) <
and d ( xn , b) <
whenever n > n0
2
2
Since e is arbitrary, take e = r then
r = d (a, b) d (a, xn ) + d ( xn , b)
r r
e
< + =r
Q d (a, xn ) = d ( xn , a) <
2 2
2
Which is a contradiction, Hence a = b i.e. limit is unique.
d ( xn , a) <

v Theorem
i)
A convergent sequence is bounded.
ii)
If xn x and yn y then d ( xn , yn ) d ( x, y ) .
Proof.
(i) Suppose xn x , therefore for any e > 0 there is n0 such that
" n > n0 , d ( xn , x) < e
Let a = max {d ( x1 , x), d ( x2 , x),............, d ( xn , x)} and k = max {e , a}
Then by using triangular inequality for arbitrary xi , x j ( xn )
0 d ( xi , x j ) d ( xi , x) + d ( x, x j )

k + k = 2k
Hence ( xn ) is bounded.
(ii) By using triangular inequality
d ( xn , yn ) d ( xn , x ) + d ( x, y ) + d ( y, yn )
d ( xn , yn ) - d ( x, y ) d ( xn , x ) + d ( y, yn ) 0 + 0
Next
d ( x, y ) d ( x, xn ) + d ( xn , yn ) + d ( yn , y )
d ( x, y ) - d ( xn , yn ) d ( x, xn ) + d ( yn , y ) 0 + 0
From (i) and (ii)
d ( xn , yn ) - d ( x, y ) 0 as n
Hence

lim d ( xn , yn ) = d ( x, y )
n

as n ...........(i )
as n ..........(ii )

Q.E.D

v Cauchy Sequence
A sequence ( xn ) in a metric space ( X , d ) is called Cauchy if any e > 0 there is a
n0 such that " m, n > n0 , d ( xm , xn ) < e .
Available at http://www.MathCity.org

13

Metric Spaces

v Theorem
A convergent sequence in a metric space ( X , d ) is Cauchy.
Proof.
Let xn x X , therefore any e > 0 there is n0 such that
e
e
" m, n > n0 , d ( xn , x) <
and d ( xm , x) < .
2
2
Then by using triangular inequality
d ( xm , xn ) d ( xm , x ) + d ( x, xn )
d ( xm , x ) + d ( xn , x )
Q d ( x, y ) = d ( y , x )
e e
< + =e
2 2
Thus every convergent sequence in a metric space is Cauchy.
v Example
Let ( xn ) be a sequence in the discrete space ( X , d ) . If ( xn ) be a Cauchy sequence, then
for e = 12 , there is a natural number n0 depending on e such that
d ( xm , xn ) < 12
" m, n n0
Since in discrete space d is either 0 or 1 therefore d ( xm , xn ) = 0 xm = xn = x (say)
Thus a Cauchy sequence in ( X , d ) become constant after a finite number of terms,
i.e.

( xn ) = ( x1, x2 ,..., xn , x, x, x,...)


0

v Subsequence
Let (a1 , a2 , a3 ,...) be a sequence ( X , d ) and let (i1 , i2 , i3 ,...) be a sequence of
positive integers such that i1 < i2 < i3 < ... then

( an : n ) .

( a , a , a ,...) is called subsequence of


i1

i2

i3

v Theorem
(i) Let ( xn ) be a Cauchy sequence in ( X , d ) , then ( xn ) converges to a point x X if

( )

and only if ( xn ) has a convergent subsequence xnk which converges to x X .

( )

(ii) If ( xn ) converges to x X , then every subsequence xnk also converges to x X .


Proof.
(i) Suppose xn x X then ( xn ) itself is a subsequence which converges to x X .

( )

Conversely, assume that xnk is a subsequence of ( xn ) which converges to x.

Then for any e > 0 there is n0 such that " nk > n0 , d xnk , x <
Further more ( xn ) is Cauchy sequence

e
.
2

e
.
2
Suppose n2 = max(n0 , n1 ) then by using the triangular inequality we have
Then for the e > 0 there is n1 such that " m, n > n1 , d ( xm , xn ) <

Available at http://www.MathCity.org

14

Metric Spaces

) (

d ( xn , x ) d xn , xnk + d xnk , x
<

e e
+ =e
2 2

)
" nk , n > n2

This show that xn x .


(ii) xn x implies for any e > 0 $ n0 such that d ( xn , x ) < e

Then in particular d xnk , x < e

" nk > n0

Hence xnk x X .
v Example
Let X = ( 0,1) then ( xn ) = ( x1 , x2 , x3 ,...) = ( 1 2 , 1 3 , 1 4 ,...) is a sequence in X.
Then xn 0 but 0 is not a point of X.
v Theorem
Let ( X , d ) be a metric space and M X .
(i)
Then x M if and only if there is a sequence ( xn ) in M such that xn x .
(ii) If for any sequence ( xn ) in M, xn x x M , then M is closed.
Proof.
(i) Suppose x M = M M
If x M , then there is a sequence ( x, x, x,...) in M which converges to x.
If x M , then x M i.e. x is an accumulation point of M, therefore each n the
1
open ball B x ; contain infinite number of point of M.
n
1
We choose xn M from each B x ;
n
1
Then we obtain a sequence ( xn ) of points of M and since 0 as n .
n
Then xn x as n .
Conversely, suppose ( xn ) such that xn x .
We prove x M
If x M then x M .
Q M = M M
If x M , then every neighbourhood of x contain infinite number of terms of ( xn ) .
Then x is a limit point of M i.e. x M
Hence x M = M M .
(ii) If ( xn ) is in M and xn x , then x M then by hypothesis M = M , then M is
closed.
v Complete Space
A metric space ( X , d ) is called complete if every Cauchy sequence in X converges
to a point of X.
Available at http://www.MathCity.org

15

Metric Spaces

v Subspace
Let ( X , d ) be a metric space and Y X then Y is called subspace if Y is itself a
metric space under the metric d.
v Theorem
A subspace of a complete metric space ( X , d ) is complete if and only if Y is closed
in X.
Proof.
Assume that Y is complete we prove Y is closed.
Let x Y then there is a sequence ( xn ) in Y such that xn x .
Since convergent sequence is a Cauchy and Y is complete then xn x Y .
Y Y

Since x was arbitrary point of Y

QY Y
Therefore Y = Y
Consequently Y is closed.
Conversely, suppose Y is closed and ( xn ) is a Cauchy sequence. Then ( xn ) is Cauchy in
X and since X is complete so xn x X .

Also x Y and Y X .
Since Y is closed i.e. Y = Y therefore x Y .
Hence Y is complete. 8
v Nested Sequence:
A sequence sets A1 , A2 , A3 ,... is called nested if A1 A2 A3 ...
v Theorem (Cantors Intersection Theorem)
A metric space ( X , d ) is complete if and only if every nested sequence of nonempty closed subset of X, whose diameter tends to zero, has a non-empty intersection.
Proof.
Suppose ( X , d ) is complete and let A1 A2 A3 ... be a nested sequence of
closed subsets of X.
Since Ai is non-empty we choose a point an from each An . And then we will prove
(a1 , a2 , a3 ,...) is Cauchy in X.

( )

Let e > 0 be given, since lim d ( An ) = 0 then there is n0 such that d An0 < 0
n

Then for m, n > n0 , d (am , an ) < e .


This shows that (an ) is Cauchy in X.
Since X is complete so an p X (say)
We prove p I An ,
n

Suppose the contrary that p I An then $ a k such that p Ak .


n

Since Ak is closed, d ( p, Ak ) = d > 0 .

Available at http://www.MathCity.org

16

Metric Spaces

d
d
Consider the open ball B p ; then Ak and B p ; are disjoint
2
2
d
Now ak , ak +1 , ak + 2 ,... all belong to Ak then all these points do not belong to B p ;
2
This is a contradiction as p is the limit point of ( an ) .
Hence p I An .
n

Conversely, assume that every nested sequence of closed subset of X has a non-empty
intersection. Let ( xn ) be Cauchy in X, where ( xn ) = ( x1 , x2 , x3 ,...)
Consider the sets
A1 = { x1 , x2 , x3 ,...}
A2 = { x2 , x3 , x4 ,...}

Ak = { xn : n k}
Then we have A1 A2 A3 ...
We prove lim d ( An ) = 0
n

Since ( xn ) is Cauchy, therefore $ n0 such that


" m, n > n0 , d ( xm , xn ) < e , i.e.

( )

lim d ( An ) = 0 .
n

Now d An = d ( An ) then lim d ( An ) = lim d ( An ) = 0


n

Also A1 A2 A3 ...
Then by hypothesis

IA

We prove xn p X

j . Let p I An
n

( )

Since lim d ( An ) = 0 therefore $ k0 such that d Ak0 < e


n

Then for n > k0 , xn , p An d ( xn , p ) < e " n > k0


This proves that xn p X .
The proof is complete.
v Complete Space (Examples)
(i) The discrete space is complete.
Since in discrete space a Cauchy sequence becomes constant after finite terms
i.e. ( xn ) is Cauchy in discrete space if it is of the form
( x1 , x2 , x3 ,..., xn = b, b, b,...)
(ii) The set = {0, 1, 2,...} of integers with usual metric is complete.
(iii) The set of rational numbers with usual metric is not complete.
Available at http://www.MathCity.org

17

Metric Spaces

Q (1.1,1.41,1.412,...) is a Cauchy sequence of rational numbers but its limit is


which is not rational.

2,

(iv) The space of irrational number with usual metric is not complete.
We take ( -1,1) , ( - 1 2 , 1 2 ) , ( - 13 , 13 ) ,..., ( - 1 n , 1 n )
We choose one irrational number from each interval and these irrational tends to zero as
we goes toward infinity, as zero is a rational so space of irrational is not complete.
v Theorem
The real line is complete.
Proof.
Let ( xn ) be any Cauchy sequence of real numbers.
We first prove that ( xn ) is bounded.
Let e = 1 > 0 then $ n0 such that " m, n n0 , d ( xm , xn ) = xm - xn < 1
In particular for n n0 we have

xn0 - xn 1 xn0 - 1 xn xn0 + 1

Let a = max x1 , x2 ,..., xn0 + 1 and b = min x1 , x2 ,..., xn0 - 1

then b xn a " n .
this shows that ( xn ) is bounded with b as lower bound and a as upper bound.
Secondly we prove ( xn ) has convergent subsequence ( xni ) .

If the range of the sequence is { xn } = { x1 , x2 , x3 ,...} is finite, then one of the term is the
sequence say b will repeat infinitely i.e. b, b ,b ,.
Then (b, b, b,...) is a convergent subsequence which converges to b.
If the range is infinite then by the Bolzano Weirestrass theorem, the bounded infinite set
{ xn } has a limit point, say b.
Then each of the open interval S1 = (b - 1, b + 1) , S2 = ( b - 1 2 , b + 1 2 ) ,
S2 = ( b - 13 , b + 13 ) , has an infinite numbers of points of the set { xn } .
i.e. there are infinite numbers of terms of the sequence ( xn ) in every open interval Sn .
We choose a point xi1 from S1 , then we choose a point xi2 from S2 such that i1 < i2
i.e. the terms xi2 comes after xi1 in the original sequence ( xn ) . Then we choose a term xi3
such that i2 < i3 , continuing in this manner we obtain a subsequence

( x ) = ( x , x , x ,K) .
in

i1

i2

i3

It is always possible to choose a term because every interval contain an infinite numbers
of terms of the sequence ( xn ) .
Since b - 1n b and b + 1n b as n . Hence we have convergent subsequence

( x ) whose limit is
in

b.

Lastly we prove that xn b .


Since ( xn ) is a Cauchy therefore for any e > 0 there is n0 such that
Available at http://www.MathCity.org

18

Metric Spaces

e
2
Also since xin b there is a natural number im such that im > n0
Then " m, n, im > n0
xm - xn <

" m, n > n0

d ( xn , b ) = xn - b = xn - xim + xim - b

xn - xim + xim - b <


Hence xn b and the proof is complete.

e e
+ =e
2 2

v Theorem
The Euclidean space n is complete.
Proof.
Let ( xm ) be any Cauchy sequence in n .
Then for any e > 0 , there is n0 such that " m, r > n0
1

( m ) ( r ) 2 2
d ( xm , xr ) = x j - x j < e ..............(i)

(m)
(r )
(m) (m) (m) (m)

(r ) (r ) (r ) (r )

where xm = x j = x1 , x 2 , x3 ,K, x n and xr = x j = x1 , x 2 , x3 ,K, x n


Squaring both sided of (i) we obtain


2

(m) (r )
x j - x j < e 2

(m)

(r )

xj-xj <e

" j = 1, 2,3,K, n

( m ) (1) (2) (3)


This implies x j = x j , x j , x j ,K is a Cauchy sequence of real numbers for every

j = 1, 2,3,K, n .
(m)

Since is complete therefore x j x j (say)


Using these n limits we define
x = (x j ) = (x1 , x 2 , x 3 ,K, x n ) then clearly x n .
We prove xm x

In (i) as r , d ( xm , x ) < e " m > n0 which show that xm x n


And the proof is complete.
Note: In the above theorem if we take n = 2 then we see complex plane = 2 is
complete. Moreover the unitary space n is complete.
v Theorem
The space l is complete.
Proof.
Available at http://www.MathCity.org

19

Metric Spaces

Let ( xm ) be any Cauchy sequence in l .


Then for any e > 0 there is n0 > such that " m, n > n0
(m)

d ( xm , xn ) = sup x j - x j < e .......... (i)


j

(m) (m) (m) (m)


(n) (n) (n) (n)
Where xm = x j = x1 , x 2 , x3 ,K and xn = x j = x1 , x 2 , x3 ,K

Then from (i)


(m)

(n)

" j = 1, 2,3,K and " m, n > n0

x j - x j < e ..........(ii)

( m ) (1) (2) (3)


It means x j = x j , x j , x j ,K is a Cauchy sequence of real or complex numbers for

every j = 1, 2,3,K
(m)

And since and are complete therefore x j x j or (say).

Using these infinitely many limits we define x = (x j ) = (x1 , x 2 , x3 ,K) .


We prove x l and xm x .
(m)

In (i) as n we obtain x j - x j < e ..........(iii ) " m > n0


We prove x is bounded.
By using the triangular inequality
(m)

(m)

(m)

(m)

x j = x j - x j + x j x j - x j + x j < e + km
(m)

Where x j < km as xm is bounded.


Hence (x j ) = x is bounded.
This shows that xn x l .
And the proof is complete.
v Theorem
The space C of all convergent sequence of complex number is complete.
Note: It is subspace of l .
Proof.
First we prove C is closed in l .
Let x = (x j ) C , then there is a sequence ( xn ) in C such that xn x ,
(n) (n) (n) (n)
where xn = x j = x1 , x 2 , x3 ,K .

Then for any e > 0 , there is n0 such that " n n0


(n)

d ( xn , x ) = sup x j - x j <
j

e
3
Available at http://www.MathCity.org

20

Metric Spaces

Then in particular for n = n0 and " j = 1,2,3,...........


( n0 )
e
x j -xj <
3
Now xn0 C then xn0 is a convergent sequence therefore $ n1

such that " j , k > n1

e
3
Then by using triangular inequality we have
( n0 )

( n0 )

x j - xk <
( n0 )

( n0 )

( n0 )

( n0 )

x j - xk = x j - x j + x j - xk + xk - xk
( n0 )

( n0 )

( n0 )

( n0 )

x j - x j + x j - xk + xk - xk

e e e
+ + =e
" j , k > n1
3 3 3
Hence x is Cauchy in l and x is convergent
Therefore x C and C = C .
i.e. C is closed in l and l is complete.
Since we know that a subspace of complete space is complete if and only if it is closed
in the space.
Consequently C is complete.
<

v Theorem
The space l p , p 1 is a real number, is complete.
Proof.
Let ( xn ) be any Cauchy sequence in l p .
Then for every e > 0 , there is n0 such that " m, n > n0
1

(m) (n)
d ( xm , xn ) = x j - x j
j =1

(m) (m) (m) (m)


where xm = x j = x1 , x 2 , x3 ,K

(m)

Then from (i)

p
< e .. (i)

(n)

x j - x j < e (ii)

" m, n > n0 and for any fixed j.

(m)
This shows that x j is a Cauchy sequence of numbers for the fixed j.

(m)

Since and are complete therefore x j x j or (say) as m .

Using these infinite many limits we define x = (x j ) = (x1 , x 2 , x3 ,K) .


We prove x l p and xm x as m .
From (i) we have

Available at http://www.MathCity.org

21

Metric Spaces
1

k (m) (n)
x j - x j
j =1
k

i.e.

(m)

x
j =1

(n) p

-x j

p
< e

< e p . (iii)

Taking as n , we get
k

(m)

x
j =1

-x j

<ep ,

k = 1, 2, 3, .

Now taking k , we obtain


(m)

-x j

< e p (iv)

" j = 1,2,3,..........

This shows that ( xm - x ) l p


Now l p is a vector space and xm l p , x - xm l p then xm + ( x - xm ) = x l p .
Also from (iv) we see that
p
" m > n0
( d ( xm , x) ) < e p
i.e. d ( xm , x) < e
" m > n0
This shows that xm x l p as x .
And the proof is complete.
v Theorem
The space C[a, b] is complete.
Proof.
Let ( xn ) be a Cauchy sequence in C[a, b].
Therefore for every e > 0 , there is n0 such that " m, n > n0
d ( xm , xn ) = max xm (t ) - xn (t ) < e (i)
Then for any fix t = t0 J

tJ

where J = [ a, b] .

xm (t0 ) - xn (t0 ) < e


" m, n > n0
It means ( x1 (t0 ), x2 (t0 ), x3 (t0 ),K) is a Cauchy sequence of real numbers. And since is
complete therefore xm (t0 ) x(t0 ) (say) as m .
In this way for every t J , we can associate a unique real number x(t ) with xn (t ) .
This defines a function x(t ) on J.
We prove x(t ) C[a, b] and xm (t ) x(t ) as m .
From (i) we see that
xm (t ) - xn (t ) < e for every t J and " m, n > n0 .
Letting n , we obtain for all t J
xm (t ) - x(t ) < e " m < n0 .
Since the convergence is uniform and the xn s are continuous, the limit function x(t ) is
continuous, as it is well known from the calculus.
Then x(t ) is continuous.
Available at http://www.MathCity.org

22

Metric Spaces

Hence x(t ) C[a, b] , also xm (t ) - x(t ) < e


Therefore xm (t ) x(t ) C[a, b] .
The proof is complete.

as m

v Theorem
If ( X , d1 ) and (Y , d 2 ) are complete then X Y is complete.

Note: The metric d (say) on X Y is defined as d ( x, y ) = max ( d1 (x1 ,x 2 ) , d 2 (h1 ,h2 ) )

where x = (x1 ,h1 ) , y = (x 2 ,h2 ) and x1 ,x 2 X , h1 ,h2 Y .


Proof.
Let ( xn ) be a Cauchy sequence in X Y .
Then for any e > 0 , there is n0 such that " m, n > n0

(m) (n) (m) (n)


d ( xm , xn ) = max d1 x , x , d 2 h , h < e


(m) (n)
(m) (n)

d1 x , x < e and d 2 h , h < e " m, n > n0

(m)
(1) (2) (3)

This implies x = x , x , x ,K is a Cauchy sequence in X.


( m ) (1) (2) (3)


and h = h , h , h ,K is a Cauchy sequence in Y.

(m)

(m)

Since X and Y are complete therefore x x X (say) and h h Y (say)


Let x = (x , m ) then x X Y .
(m) (m)
d ( xm , x ) = max d1 x ,x , d 2 h ,h 0 as n .


Hence xm x X Y .
This proves completeness of X Y .

Also

v Theorem
f : ( X , d ) (Y , d ) is continuous at x0 X if and only if xn x implies
f ( xn ) f ( x0 ) .
Proof.
Assume that f is continuous at x0 X then for given e > 0 there is a d > 0
such that
d ( x, x0 ) < d d ( f ( x), f ( x0 ) ) < e .
Let xn x0 , then for our d > 0 there is n0 such that
d ( xn , x0 ) < d , " n > n0
Then by hypothesis d ( f ( xn ), f ( x0 ) ) < e , " n > n0
i.e.
f ( xn ) f ( x0 )
Conversely, assume that xn x0 f ( xn ) f ( x0 )
We prove f : X Y is continuous at x0 X , suppose this is false
Available at http://www.MathCity.org

23

Metric Spaces

Then there is an e > 0 such that for every d > 0 there is an x X such that
d ( x, x0 ) < d but d ( f ( x), f ( x0 ) ) e
1
In particular when d = , there is xn X such that
n
d ( xn , x0 ) < d but d ( f ( xn ), f ( x0 ) ) e .
This shows that xn x0 but f ( xn ) f ( x0 ) as n .
This is a contradiction.
Consequently f : X Y is continuous at x0 X .
The proof is complete.
v Rare (or nowhere dense in X )
Let X be a metric, a subset M X is called rare (or nowhere dense in X ) if M
has no interior point i.e. int M = j .

( )

v Meager ( or of the first category)


Let X be a metric, a subset M X is called meager (or of the first category) if M
can be expressed as a union of countably many rare subset of X.
v Non-meager ( or of the second category)
Let X be a metric, a subset M X is called non-meager (or of the second category)
if it is not meager (of the first category) in X.
v Example:
Consider the set of rationales as a subset of a real line . Let q , then {q} = {q}
because - {q} = ( -, q ) ( q, ) is open. Clearly {q} contain no open ball. Hence is
nowhere dense in as well as in . Also since is countable, it is the countable union
of subsets {q} , q . Thus is of the first category.
v Bairs Category Theorem
If X j is complete then it is non-meager in itself.
OR
A complete metric space is of second category.
Proof.

Suppose that X is meager in itself then X = U M k , where each M k is rare in X.


k =1

Since M 1 is rare then int( M ) = M = j


i.e. M 1 has non-empty open subset
But X has a non-empty open subset ( i.e. X itself ) then M 1 X .
c

This implies M 1 = X - M 1 is a non-empty and open.


We choose a point p1 M 1

c
and an open ball B1 = B ( p1; e1 ) M 1 , where e1 < 1 .
2

Available at http://www.MathCity.org

24

Metric Spaces
c

Now M 2 is non-empty and open


c
c
1
Then $ a point p2 M 2 and open ball B2 = B ( p2 ; e 2 ) M 2 B p1; e1
2
c
1
( M 2 has no non-empty open subset then M 2 B p1; e1 is non-empty and open.)
2
c
1
So we have chosen a point p2 from the set M 2 B p1; e1 and an open ball
2
1
1 1
B ( p2 , e 2 ) around it, where e 2 < e1 < < 2-1 .
2
2 2
Proceeding in this way we obtain a sequence of balls Bk such that
1

Bk +1 B pk ; e k Bk where Bk = B ( pk ; e k ) " k = 1, 2,3,.......


2

Then the sequence of centres pk is such that for m > n


1
1
d ( pm , pn ) < e m < m +1 0 as m .
2
2
Hence the sequence ( pk ) is Cauchy.
Since X is complete therefore pk p X (say) as k .
Also
d ( pm , p ) d ( pm , pn ) + d ( pn , p )
1
< e m + d ( pn , p )
2
< e m + d ( pn , p ) e m + 0 as n .

p Bm

" m

i.e. p M m

" m

Q Bm = M 2 B pm-1; 12 e m-1
c

Bm M m
Bm M m = j
p Mm " m p X
This is a contradiction.
Bairs Theorem is proof.

References: (1) Lectures (2003-04)


Prof. Muhammad Ashfaq
Ex Chairman, Department of Mathematics.
University of Sargodha, Sargodha.
(2) Book
Introductory Functional Analysis with Applications
By Erwin Kreyszig (John Wiley & Sons. Inc., 1989.)

These notes are available online at http://www.mathcity.org in PDF Format.


Last update: July 01, 2011.

Available at http://www.MathCity.org

Anda mungkin juga menyukai