You are on page 1of 14

Chapter 6

SEQUENCES
6.1 Sequences
A sequence is a function whose domain is the set of natural numbers N. If s is a

sequence, we usually denote its value at n by sn


sequence

as

instead of

{sn } or by listing the elements {s1 , s2 , ....} .

s (n) . We may

refer to the

This notation is somewhat

ambiguous though, since a sequence is an ordered set, and should not be confused with

s = {sn : n N} . Some authors prefer the use of parentheses instead


of braces, to emphasize this particular feature of sequences. Our notation will use
braces for both.
We call sn the nth term of the sequence and we often describe a sequence by
a formula
for the nth term. Thus {1/n} is an abbreviation for the sequence
giving

1 1
1, 2 , 3 , .... . Sometimes we may wish to change the domain of a sequence from N to
N{0} , or to {n N : n m} . In this case we write {sn }n=0 or {sn }n=m respectively.
the range of

N.

If no mention is made to the contrary, we assume that the domain is just

Sequences are sometimes also expressed in what we call recursive form. This

sn is not expressed explicitly as a function of n, but


instead
as a function ofsome subset of previous terms {sn1, sn2 ..., snk } . Thus

 1 1 
s 1
sn : sn = snn
,
s
= 1
is again an abbreviation for the sequence 1, , ... Note
1
+1
2 3
1
that whenever the recursive definition for sn depends on k previous terms, for the
sequence to be well defined it is necessary to provide the first k terms .
means that a generic term

6.2 Convergence

6.2.1 Convergent sequences


{sn } converges to a certain value s whenever
the sn become arbitrarily close to s for n arbitrarily large. It is clear from the above
that we need a metric space to define what close means. In this section we will
restrict to the metric space ( , ||) . The formal definition of convergent sequence in
an arbitrary metric space is the following:

Intuitively, we say that a given sequence

Definition 224 A sequence


sX
there exists

for all

nN

{sn }

in a metric space (X,d)

such that for every

and

>0

n N , d (sn , s) < .

is said to

there exists a real number

converge

if

such that

76

Sequences

{sn } converges to s, or that s is the limit of the


sn s, lim sn = s, or limn sn = s. If a sequence does

In this case we say that

sequence {sn } ,

and write

not converge, it is said to diverge.

Remark 225 Notice that for a sequence to be convergent the definition requires its
limit to belong to X, the set in which the sequence is defined. Take for example the
sequence {1/n}. Even though it converges
to 0 in X = R, it fails

 to do so in X= R++ .
xn + 2
In the same fashion, the sequence x
x
, x = 1 converges to 2 in R
n

n+1

2 xn

Q.
Example 226 The sequence {sn} with sn = 1 for all n N converges to 1 in R.
,but it fails to converge in
The proof is trivial since

d(sn , 1) = 0 < for all sn

Example 227 Let us show that


we have to show that for all > 0,

and for all

0.

. According to the definition,


M R, s.t. n N, n  M, 1/n 0 =
1/n < . So take an arbitrary > 0, R. From the Archimedean property of R
there exits M N R such that M > 1/. Then it follows that for all n M we
have 1/n  1/M < , which concludes the proof.
Example 228 To prove that the sequence {1 + (1)n} is divergent, let us suppose
that {sn} converges to some real number s. Letting = 1 in the definition of converlim 1

/n

R,

>

||)

= 0 in (

n
gence, we find that there exists a number M such that |1 + (1) s| < 1. If n > M
n
and n is odd, then |1 + (1) s| = |s| < 1, so that 1 < s < 1. On the other hand, if
n
n > M and n is even, then we have |1 + (1) s| = |2 s| < 1, so that 1 < s < 3.
Since s cannot satisfy both inequalities, we have reached a contradiction. Thus the
sequence {sn } must be divergent.

So far we have seen examples where algebraic manipulation was enough to

find an and establish whether the definition was satisfied or not. However, for
more complicated cases such a simple technique may not work.

Example 229 Take for example the case of


lim

sn

= 0

From the definition, given any >

We want to show that


n
, we want to make nn +2
5 < . By

sn

2
n
= nn3+2
5 .


 2
 3





n > 0 . Thus we want to


restricting n > 3,we can remove the absolute value since nn23+2
5
know how big n has to be in order to make nn23+25n < . Solving for this inequality would

be very messy, so we try to find some estimate of how large the left hand side can
be. To do this we seek an upper bound for the numerator and a lower bound for the
denominator. On the one hand we have n2 + 2n < a n2 for some a > 1 if n > (a2 1) .

On the other we have n3 5 > nb

for some

pair a, b > 1 and restricting n > max 3, (a2 1)


n + 2n  n2 + 2n
=
n 5  n3 5


 2

 3

 

b>1
3

n>

if

5 bb 1


3

5 bb 1 .

we have that

ann = abn .
2

Taking an arbitrary

Convergence

77

The right hand side can be easily made less than any . Just pick n ab . So
now we are ready
this into a formal proof. Given > 0, take M =
3 bto organize

2
ab
max 3, (a1) 5 b1 , . Then for all n N and n > M we have
,






Hence

lim


n2 + 2n
n2 + 2n

0 = n3 5
n3 5


a n2
n3

ab
<
n

= 0.

n2 +2n
n3 5

The use of upper bounds can be generalized by the following theorem

Exercise 230

Suppose that

lim sn = 0.

If {tn }

lim (sn tn ) = 0. Use this result to prove that lim

Theorem 231

is a bounded sequence, prove that

sin(n )

=0

Let {sn } and {an } be sequences of real numbers and let s R . If for

some k > 0 and some m N we have

|sn s| k |an | ,

for all

and if lim an = 0 , then it follows that lim sn =

Proof. Given any > , since


0

lim

n > N1 implies that |an | < /k. Now


have n > m and n > N1, so that

and

an

s.

= 0

lim

sn

there exists N1 R such that n N

let N = max

|sn s| k |an | < k

Thus

n>m

m, N1} .

Then for n > N we

s.

Another useful theorem to show convergence of sequences by use of bounding


sequences is the following.

Theorem 232 (Sandwich theorem) Suppose xn  yn  zn


M

is fixed, limn

for

xn = l, and limn zn = l, then limn yn = l.

n > M,

where

Exercise 233 Prove the Sandwich theorem.


There is also a necessary condition for convergence that is very useful both
in proofs of convergence and in proving that a sequence does not converge. It states
that for n large enough, any two terms in a sequence must be arbitrarily close to each
other.

78

Sequences

Theorem 234

that for all

If a sequence

> 0 there is N R such

{sn } converges, then for all

n1 , n2 N , n1 , n2 > N , |sn1 sn2 | < .

Proof.

n N,n > N

Since {sn } converges, for any > 0 there is N R

we have

such that for all

|sn s| < 2 . In particular, if we take n1 > N

and

n2 > N

we

have by the triangle inequality

sn1 sn2 | |sn1 s| + |s sn2 | <

+
2 2 = .

Exercise 235 Prove that the reciprocal may not be true. That is, prove that you can

have a sequence whose terms are arbitrarily close for n sufficiently large, but still the
sequence does not converge.
6.2.2 Properties of Convergent Sequences
In this section we derive two important properties of convergent sequences. But first
we need a definition.

Definition 236 A sequence {sn } is bounded if its range {sn

n N}is

|sn | M for all n N.


Theorem 237 (Boundedness) Every convergent sequence is bounded.
Proof. Let {sn} be a convergent sequence and let s = lim sn. From the definition of convergence with = 1, we obtain N R such that |sn s| < 1 whenever
n > N. Thus for n > N the triangle inequality implies that |sn | < |s| + 1. Let
set, that is, if there exists

= max

M 0

a bounded

such that

s1| , . . . , |sN | , |s| + 1},

{|

then we have |sn |

for all

N, so

sn }

is

bounded.

It is easy to see that the converse of this theorem is not true. We saw in a
previous example the sequence

{1 + (1)n },

which is bounded but diverges.

Theorem 238 (Uniqueness) If a sequence {sn} converges, then its limit is unique.
Proof. Let {sn} be a sequence and suppose that it converges both to x and y,
with x = y . Let x y = d > 0 be the distance between x and y. Convergence of
sn to x implies that for any > 0, there exists N1 R such that
|sn x| < ,
n > N1.
= d3 Similarly, convergence of {sn } to y implies that for
|

for every

In particular this holds for

any

> 0,

there exists

N2

R such that

|sn y| < ,

for every

n > N2,

= d3 . Therefore if n > max {N1, N2} we have


d
0 < d = |x y| = |x sn + sn y | |x sn | + |sn y | < 2
3

and in particular this holds for

But this implies


d = 0,

and thus

d < 23 d,
x = y.

which is a contradiction given that

d > 0.

So it must be

Convergence

79

6.2.3 Convergence and topological properties of sets


In the previous chapter, we saw that a set

X was closed if and only if

it contained all

its accumulation points. Here we will present the same result based on the theory of

xn x, then x is an accumulation point for the set described

sequences. But in order to do that, we have first to prove two other theorems. The
first one states that if

by the range of the sequence, and hence every neighborhood of


many points of

{xn} .

contains infinitely

The second theorem establishes that it is equivalent to say that

E
in E.

a point belongs to the closure of a given set


limit of a sequence completely contained

and to say that such a point is the

{xn } X converges to x X if and only if every neighborhood of x


{xn } for all but finitely many n.
Proof. If {xn} converges to x then for each > 0 there exists N R such
that for all n N, n > N d (xn , x) < . This means that any neighborhood of x of
Theorem 239
contains

radius has at least one (hence infinitely many) elements of {xn}. Hence there must
be at most only finitely many elements of {xn} outside any neighborhood of x.
Suppose now that every neighborhood of x contains {xn} for all but finitely
many n. Hence for each neighborhood N (x, ), N R such that for all n N,
n > N, xn N (x, ). This suffices to conclude that {xn } converges to x.
Theorem 240 Let E X . Then x E if and only if there exists a sequence {xn}
in E such that x
n xn .
= lim

Proof. Suppose there exists a sequence {xn} E that is convergent to x X


Therefore x E
C . Since E
C is open, there exists a neighborhood N (x, )
but x
/ E.
E C such that N (x, ) {xn : n N} = . Therefore, for such there is no N R such
that for all n > N |xn x| < , contradicting the assumption that x = limn xn.

Then x E.
Then x E or x bd E. This implies that for every
Suppose now that x E.
> 0, N (x, ) E = . We cantherefore construct a sequence {xn} in E , being xn
any point such that xn N x, n1 E. By the Archimedean property of the reals, for
all > 0 there exists N N such that 1/N < , and by construction of the sequence
there exists xN {xn } such that |xN x| < 1/N < . Therefore x = limn xn.

We have therefore seen that any limit point of E can be expressed as a limit
of a convergent sequence in E. So, this theorem tells us that any nonempty closed
set contains convergent sequences. Moreover, we know that if in a given set E every
convergent sequence has a limit in the set, then the set is closed. Therefore we can
define the closedness of a set in a metric space using convergent sequences.
Theorem 241 A set E X is closed if and only if every convergent sequence in X

completely contained in E has its limit in E.


Proof. Let E be a closed set and let {xn} be a sequence in E converging to
Since E is closed, then E = E

x X. By the previous theorem we know that x E.


and therefore x E.

80

Sequences

Suppose now that every convergent sequence in X completely contained in E


has its limit in E, but E is not closed. Then there exists x E E and x
/ E. As


in the previous theorem there exists a sequence {xn } E such that xn x E E.


But this contradicts that every convergent sequence {xn } E must have its limit in
E . Therefore E is closed.

6.3 Limit and comparison theorems


In the previous section, we saw that the definition of convergence may sometimes be
messy to use even for sequences given by relatively simple formulas. In this section
we will derive some basic results that will greatly simplify our work.
The first theorem is a very important result showing that algebraic operations
are compatible with taking limits.

Theorem 242 Consider two sequences {xn} and {yn} such that xn x and yn y.
Then the following properties hold:
1) xn + yn x + y ;
2) xn yn xy ;
3) If xn =
 0, n and x = 0, then yn /xn y/x.
Proof. 1) Using the triangular inequality, we know that |xn + yn x y| 
|xn x| + |yn y|. Now given > 0, since xn x there exists N1 R such

n > N1 implies |xn x| < 2 Similarly there exists such N2 R such that
n > N2 implies |yn y | < 2 . Thus, if we let N = max{N1 , N2 }, then n > N implies
.

that

that

| xn + yn ) (x + y )|  |xn x| + |yn y| < .

Therefore we conclude that


xn yn
x y.
2) Observe that xnyn xy xn yn y y xn x . Since {xn } is convergent
then it is bounded. Let M1 = sup xn and M = max {M1, |y|}. Hence
lim (

) =

)+

| n yn xy|  M1 |yn y| + |y| |xn x| M

Now given > 0 there exist Nx and Ny such that

for all n > Nx and n > Ny respectively. Let N

that

Thus
That

|xn yn xy|

yn y| + |xn x| ) .

and |y y | <
xn x| < 2M
n
2M
{ x y }. Then
implies

= max N , N

n>N

M ( yn y + xn x )
< M 2M + M 2M = .


xy.
3) Since yn /xn = yn (1/xn), it suffices from (2) to show that lim (1/xn) = 1/x.
is, given > 0, we must make
lim

xn yn

x xn 

=
<

xn x
xn x 


 1



1 






Limit and comparison theorems


for

n sufficiently large.

x = 0,

there exists

have

81

To get a lower bound on the denominator, we note that since

N1 such that n > N1 implies xn x| <


|

x xn | <

| |

x|

Thus for n > N1 we

|x| |xn | < |x2| |x2| < |xn|

and

x xn  |xn x|
=
<
xn x  |xn | |x|

2 |xn x|
|x|2
But since xn x, for every > 0 there exists also N2 such that for n > N2 we have
xn x| < x2 . Let N = max {N1, N2} . Then n > N implies


 x xn 
2

<
x x <
 x x 
|x|2 n
n





| |

Hence
is

lim 1

axn

/xn = 1/x.

ax, a R

when

xn

a + xn

x .

A particular case of (1) is

x, a

Another useful fact is that the order relation


limits. The proof is left as an exercise.

Theorem 243
and

lim

zn

6.3.1

xn

x ,

and of (3)

is preserved when taking

Suppose that {yn } and {zn } are convergent sequences with lim yn = s

z. If yn

Exercise 244

when

zn for all n > M, for certain M N , then y z.

Prove the previous theorem.

Infinite Limits

The sequence sn = n is clearly not convergent, since it is not bounded. But its
behavior is not the least erratic: the terms get larger and larger. Although there is
no real number that the terms approach, we would like to say that sn goes to .
We make this precise in the following definition.

Definition 245 A sequence {xn} of R diverges to M > , N >

xn = +. A sequence
{xn } diverges to if M > 0, N > 0 such that n > N, xn < M. We write
xn or limn xn = .
that

n > N, xn > M.

We write

xn
+

if

0 such

or limn

The technique of developing proofs for infinite limits is similar to that for finite
limits, and is closely related to theorem243. We leave the proof as an exercise.

82

Sequences

Theorem 246

Suppose that {sn } and {tn } are sequences such that sn tn for all

n > M, for a certain M N.


(a) If lim sn = + , then
(b) If lim tn =
, then

lim
lim

.
.

tn = +
sn =

Exercise 247 Prove the previous theorem.


Exercise 248
Prove the
following

1)

lim

n+1 n

4
2) lim n2n+13
5 +3 =
n
!
3) lim nn = 0

= 0

6.4 Monotone sequences and Cauchy sequences


In the preceding two sections, we have seen a number of results that enable us to
show that a sequence converges. Unfortunately, most of these techniques depend on
our knowing what the limit of the sequence is before we begin. Often in applications
it is desirable to be able to show that a given sequence is convergent without knowing
precisely the value of the limit. In this section we obtain two important theorems
that do just that.

6.4.1 Monotone Sequences


Definition 249 A sequence {xn} of real number is increasing if xn+1  xn (strictly

xn+1 > xn ) n
xn+1 < xn ), n N and
increasing if

N; decreasing

monotone

if

xn+1

 xn (strictly decreasing if

if it is increasing or decreasing (strictly mono-

tone if it is strictly increasing or strictly decreasing).

Example 250

The sequence of real numbers

and bounded. The constant sequence

{1/n}

is decreasing (thus monotone)

x1 = x2 = ... = 1

is increasing, decreasing, and

bounded. Both of them converge.

Example 251 The sequence {(1)n } is bounded, but it is not monotone. The sequence {2n } is monotone but it is not bounded. None of them converge.
As you may have imagined, we can generalize these examples in a theorem
Theorem 252 (Monotone Convergence Theorem) A monotone sequence of real
numbers {xn} converges if and only if it is bounded.

Proof. We already know that a convergent sequence is bounded. Hence we just


need to show that a monotone and bounded sequence of real numbers is convergent.
Suppose {xn} is a bounded increasing sequence. Let S denote the nonempty bounded
set {xn n N}. By the completeness axiom S has a least upper bound, and we let
:

x = sup S . We claim that lim xn = x. Given

any > 0, x is not an upper bound

Monotone sequences and Cauchy sequences

83

for S. Thus there exists an integer N such that

xN > x . Furthermore, since {xn }

is increasing and x is an upper bound for S we have


x < xN

xn x < x

or equivalently
for all n

N. Hence

|xn x| <
lim

xn

n. In the case when the sequence is decreasing, let

x = inf S and proceed in a similar manner.

Exercise 253 Consider the sequence


2 ,

{xn}

xn : xn+1 =

is decreasing and bounded below.

x2n +2
2xn , x1

=1

In particular, show

xn n 2 (try induction)
. Thus we know that {xn }
that {xn } converges to
(Hint: use theorem 234).

. Show that for n

that
2 < xn+1 <

is convergent in

R.

Show also

6.4.2 Cauchy sequences

When a sequence is convergent, we saw in theorem 234 that the terms get close to each
other as n gets large. It turns out that this property ( called the Cauchy property)
is actually sufficient to imply convergence.
in some cases

A sequence {xn } in a metric space (X, d) is said to be a Cauchy


sequence if for all > 0, there exists a number M s.t. for all n,m N, n, m  M
Definition 254

d(xn , xm ) < .

Example 255 The sequence {1/n} R++ is Cauchy. For each > 0 take M = 2 ,
1
implying that for all n,m N and n, m > M , |1/n 1 /m| < |1 /n| + |1 /m| < 2
M <

. Then the sequence is Cauchy. However, it is not convergent since 0 does not belong
to R++.

Exercise 256 Prove that the sequence {( ni=1 1/n)} is not Cauchy.
As we saw in theorem 234, convergence is a sufficient condition for a sequence
to be a Cauchy sequence.

Theorem 257
Proof.
Exercise 258

Every convergent sequence is a Cauchy sequence.


See theorem 234.
Show that every Cauchy sequence is bounded.

Its contrapositive If a sequence is not Cauchy, then it is not convergent is


very a useful tool to prove that a sequence does not converge. Its converse however, is
not true as we have shown in the previous example. However, there are some metric
spaces where Cauchy sequences do converge. These spaces are called complete.

84

Sequences

Definition 259 A metric space (X,d)


X
X.

in

is called

complete

if every Cauchy sequence

converges in

Theorem 260
Proof.

R, d) is complete.
Suppose that {sn } R is a Cauchy sequence and let S = {sn : n N}
The metric space

be the range of the sequence. We consider two cases, depending on whether S is finite
or infinite.
If S is finite, then the minimum distance between distinct points of S is
positive. Since {sn } is Cauchy, there exists a number N such that m, n > N implies
that |sn sm | < . Let n0 be the smallest integer greater than N. Given any m > N ,
sm and sn0 are both in S, so if the distance between them is less than , it must be
zero (since is the minimum distance between distinct points in S). Thus sm
sn 0
for all n > N. It follows that lim sn
sn0 .
Now suppose that S is infinite. From the previous exercise we know that S
is bounded. Thus from the Bolzano Wierestrass theorem, there exists a point s in

that is an accumulation point of

> 0,

S.

We claim that

{sn }

converges to

s.

Given any

N R , such that |sn sm| <  whenever



 n, m > N.
 Since s

is an accumulation point of S, the neighborhood N s, 2 = s 2 , s + 2 contains


there exists

infinitely many
of S. Thus in particular there exists an integer
 points


that m
n>N
2 .

N s,

Hence, for any

sn s|

we have

|sn sm sm s|
|sn sm | |sm s| <
=

and lim sn

implies that

s. Therefore every Cauchy sequence in

(R, d) is a complete space.

Corollary 261

m>N

such

R is a convergent sequence, which

R is convergent.
Note that in the proofs of both theorem, the completeness axiom of R is the
key factor. So, you can naturally guess that these theorems do not hold for a sequence
of Q, as it is the case.
Every Cauchy sequence in

6.5 Subsequences

Definition 262 Let {sn }


n=1 be a sequence and let {nk }k =1 be any sequence of natural numbers such that n1 < n2 < n3 < ... . The sequence {xnk }
k=1 is called a subse
quence of {xn }n=1.

When viewed as a function, a subsequence of a sequence s : N R


s
n N N.
sn
NR
s N R.

is the

composition of

and a strictly increasing function

is a subsequence of the sequence

That is,

Since this functional notation

is cumbersome, we prefer to use subscripts and write

important to note that the index of the subsequence is

sn k

instead of

k, not nk

s (n (k )) .

and not

n.

It is

Subsequences

85

If we delete a finite number of terms in a sequence and renumber the remaining

, we obtain a subsequence. In fact, we may delete infinitely


many of them as long as there are infinitely many terms left. It is possible to do so
because a countable set has countable subsets (theorem 157).
ones in the same order

Example 263
subsequence

Given the sequence {x1 , x2 , x3 , . . . , xn , . . . } the following represents a


{x1 , x3, x5 , . . . , x2n+1 , . . . } and it is obtained by taking into account only

the odd terms of the original sequence.

Exercise 264 Let {nk }k=1 be a sequence of natural numbers such that nk < nk+1 for
all k N. Use induction to show that nk k for all k N.
Theorem 265 A sequence {sn} converges to a real number s if, and only if, every
subsequence of {sn } also converges to s.

Proof. Let {sn} converge to s and take any subsequence {snk }. We know that
for each > 0 there exists a number N such that n > N implies d(sn ,s) < . Thus
when k > N, we apply the result of the previous exercise to obtain that nk k > N
and therefore d(snk , x) < .

Suppose now that every subsequence {snk } converges to s. Hence {sn } con-

verges as well because it is also a subsequence of itself.

Example 266

One application of this theorem is in finding the limit of a convergent


sequence. Suppose that 0 < x < 1 and consider the sequence defined by sn = x1/n .
Since 0 < x1/n < 1 for all n, {sn } is bounded. Since

x1/n+1 x1/n
then

{sn }

= x1/n+1 1 x nn+1

is an increasing sequence.

> 0,

for all

Thus by the monotone convergence theorem

{sn } converges to some number, say s. Now, for each n, s2n = x1/(2n) = sn . Since
{s2n } is a subsequence of {sn } then it also converges to s. Moreover, we have that
lim
sn = lim sn , and this implies

= lim

sn

= lim

s2n

= lim

sn 
=

lim

sn

s.

= s, so that s = 0 or s = 1. Since s1 = x > 0 and sn is increasing,


so s 
. Hence
x1/n = 1
Exercise267 Prove
that if {sn }
sn s

It follows that

s2

= 0

lim

is a sequence of nonnegative terms and

then lim

sn

lim

sn .

Example 268 Another application of the previous theoremn is in showing that a sequence is divergent. For example, if the sequence sn = (1) were convergent to some

number s, then every subsequence would also converge to s. But s2n } converges to 1
and {s2n+1} converges to -1. We conclude that {sn } is not convergent.

86

Sequences

So far we know that every subsequence of a convergent sequence converges


to the same limit and that in a divergent sequence there exists a subsequence that
diverges. Now the question is whether we can find a convergent subsequence in a
divergent sequence. A version of the Bolzano-Wierestrass theorem for sequences says
that if the sequence is bounded this is so.

Theorem 269 Every bounded sequence of real numbers has a convergent subsequence.
Proof. The proof is essentially the same as the Bolzano-Weierstrass theorem

we have seen in the previous chapter and is left as an exercise.

While an unbounded sequence may not have any convergent subsequence, it


will contain a subsequence that has an infinite limit. We leave the result as an
exercise.
Exercise 270

Show that every unbounded sequence contains a monotone subsequence


that has either +
or - as a limit.

6.5.1

Limit Superior and Limit Inferior


Let sn be a bounded sequence. A subsequential limit s is any
real number that is the limit of some subsequence of
{sn }. If S is the set of all
subsequential limits of {sn }, then we define the limit superior to be

Definition 271

{ }

sn

lim sup

Similarly, we define the

= sup

S.

of {sn } to be

limit inferior

sn = inf S.

lim inf

Note that in the definition we require


implies that

{sn }

to be bounded. Thus theorem 269

contains a convergent subsequence, so that the set S of subsequen-

tial limits will be nonempty.

It will also be bounded, since

completeness axiom then implies that sup

Example 272 Let sn = (

1)

and inf

{sn }

is bounded.

The

both exist as real numbers.

/n). Then

+ (1

lim sup

{sn}

{sn }

sn

= 1

sn = 1.

lim inf

It should be clear that we always have lim inf n

sn lim supn sn . Now,

s, then all its subsequences converge to s, so we


have lim inf n sn = lim supn sn . The converse is also true. If it happens that
lim inf n sn < lim sup n sn , then we say that sn
oscillates.

if

converges to some number

{ }

Subsequences

87

Theorem 273 For a real valued sequence {sn }, limn sn = s

lim inf n

sn = s.

Proof.

if and only if lim supn

sn =

sn converges to s so do all its subsequences. Hence S is singleton


S. Suppose now that sup S = inf S = s. This means that S is singleton
so that every converging subsequence of {sn } converges to s. Hence {sn } converges to
s.
If

and sup S = inf

Theorem 274

Let xn , yn be bounded sequences and xn


yn for n > M, where M is
fixed. Then lim supn xn
lim sup n yn , and lim inf n xn
lim inf n yn .
Let = lim supn xn and Y = lim supn yn and suppose that X >
Y. Hence for some {xnk } and {ynk } subsequences of {xn } and {yn } we have that for

Proof.

0
X <

nk > N = max{N,x, N,y },


< (X Y )/2 and let n = max{N,M }.

every > there exist an N,x and N,y such


|xnk |
and | nk |
. Take
Hence, for all k
we have that

y
n > n,

Y <

that for all

xnk ynk > X Y 2 > 0

which gives a contradiction, meaning that lim supn xn  lim supn yn .

The part relative to


Exercise 275 Let sn n
=

can be proved in a similar way.

lim inf

sin

 
n .
2

Find the set S of subsequential limits, the limit

superior and the limit inferior of {sn } .

Unbounded sequences

There are some occasions when we wish to generalize the notion of the limit superior and the limit inferior to apply to unbounded sequences. There are two cases
to consider for the limit superior, with analogous definitions applying to the limit
inferior.
1. Suppose that {sn} is unbounded above. Then exercise 270 implies that there exists a subsequence having + as its limit. This prompt us to define lim supn sn =

+ .

{ }

2. Suppose that sn is bounded above but not below. If some subsequence converges to a finite number, we define lim supn sn to be the supremum of the set
of subsequential limits. Essentially, this coincides with the original definition. If
no subsequence converges to a finite number, then we must have lim sn =
,
so we define lim supn sn =
.

{ }

Thus for any sequence sn , lim supn sn always exists as either a real number, + or
. When k
R and = lim supn sn , then writing > k means
that is a real number greater than k or that = + . Similarly, < k means that
is a real number less than k or that =
.

88

Sequences

6.5.2 Some special sequences

The following are very common sequences. You need to be able to prove the results
stated.

Exercise 276

Prove the following statements


1. If p > 0 then limn ( n1 )p = 0;

2. If

3.

p>0

limn

4. If

p>0

5. If

| |

n n = 1; n n p
then lim

and

x < 1,

R,

then

limn

then

= 1;

limn

xn

= 0.

(1+p)n

= 0;

6.6 References
S.R. Lay, Analysis with an Introduction to Proof .Chapter 4. Third Edition. Prentice Hall.
A. Matozzi, Lecture Notes Econ 897 University of Pennsylvania Summer
2001.
Rudin, Principles of Mathematical Analysis . Chapter 3.

You might also like