16MT AnalysisI Webnotesall
16MT AnalysisI Webnotesall
ANALYSIS I
Webnotes by H.A. Priestley
1
2
Contents
A summary of the axioms for the real numbers is given in a separate reference
sheet. References [BS ... ] are to the textbook by Bartle and Sherbert.
Introductory remarks on the content, aims and style of the Analysis I course, and how it
relates to school mathematics and to other Prelims courses.
We introduce the axioms for R and explore how the familiar rules of arithmetic can be
obtained as consequences of the axioms (and of properties derived from the axioms).
See reference sheet Axioms for the Real Numbers for the list of axioms.
Addition
The operation + of addition and the axioms, A1–A4, for addition.
Multiplication and avoiding collapse
The operation · of multiplication and the axioms, M1–M4, for multiplication. The need
to assume 0 6= 1 (Axiom Z).
Distributive Law
Linking addition and multiplication together via the distributive law D.
(10) a · 0 = 0.
(11) a · (−b) = − (a · b). In particular (−1) · a = −a.
(12) (−1) · (−1) = 1.
(13) If a · b = 0 then either a = 0 or b = 0 (or both). Moreover, if a 6= 0 and b 6= 0 then
1/(a · b) = (1/a) · (1/b).
Full proofs are given below but only a small selection of these will be given in
lectures, to illustrate how derivations from the axioms should look. You are
recommended to work through some of the others by yourself, referring to the
notes only to check that you haven’t cut corners or made correct assertions
you have omitted to validate.
Claims (6)–(8) do for multiplication what (1)–(3) do for addition, but with inverses neces-
sarily considered only for non-zero a, as M4 requires. The proofs go the same way.
(9) Note that the statement is like Axiom D but with multiplication on the other side. We
have
(a + b) · c = c · (a + b) by M1
=c·a+c·b by D
=a·c+b·c by M1 twice.
8
(10) We have
a·0+0=a·0 by A3
= a · (0 + 0) by A3
=a·0+a·0 by D.
(a · b) + (a · (−b)) = a · (b + (−b)) by D
=a·0 by A4
=0 by (10). Also
(a · b) + (−(a · b)) = 0 by A4.
This contradicts Axiom Z. The second assertion has been proved along the way.
We now have established that, according to the axioms we have set up, arith-
metic behaves as we expect it to. We shall henceforth not spell out uses of
the axioms in such detail, and are ready to revert to more familiar notation.
We may, thanks to A2 and M2, omit brackets in iterated sums and products, and write
for example a + b + c without ambiguity.
9
1.3. Powers.
Let a ∈ R \ {0}. As usual, we define a0 = 1. We then define
(
ak+1 = ak · a for k = 0, 1, 2, . . . (an ‘inductive’ or ‘recursive’ definition),
a` = 1/(a−` ) for ` = −1, −2, . . . .
Then in particular a1 = a and a2 serves (as expected) as shorthand for a·a. Problem sheet 1,
Q. 3 asks for a proof of the familiar law of indices.
1.4. Stocktaking so far, and looking further afield: R compared to other systems.
Substitute a set F in place of R in the above, and assume that the operations + and ·,
now defined on F , satisfy the axioms
A1–A4,
M1–M4,
Z,
D.
Then F , or more precisely (F ; +, ·), is a field. So our assumptions about R so far can be
summed up as
(R; +, ·) is a field.
The number systems Q (rational numbers) and C (complex numbers) are also fields. But N
(the set of natural numbers, 0, 1, 2, . . .) and Z (the integers) have weaker arithmetic properties
and are not fields. [In the Prelims Linear Algebra courses, in the definition of a vector space,
the scalars are assumed to be drawn from any field.]
2.1. Positive numbers; order relations; the real numbers as a ‘number line’.
There is a subset P (the (strictly) positive numbers) of R such that, for a, b ∈ R,
P1 a, b ∈ P =⇒ a + b ∈ P;
P2 a, b ∈ P =⇒ a · b ∈ P;
P3 exactly one of a ∈ P, a = 0 and −a ∈ P holds.
We write a < b (or b > a) iff b − a ∈ P and a 6 b (or b > a) iff b − a ∈ P ∪ {0} (the
non-negative numbers).
We subsequently make use of 6 or of <, as convenient.
[The Mean Value Theorem (from Analysis II) allows one to extend the result by replacing
n by any real number > 1.]
11
Proof. We shall prove the inequality by induction—note that the inequality is trivially true
when n = 1.
Suppose that, for k ∈ N,
(1 + x)k > 1 + kx
holds for all real x > −1. Then 1 + x > 0 by 2.3(3) and kx2 > 0 as k > 0 and x2 > 0 by
2.3(5).
(1 + x)k+1 = (1 + x) (1 + x)k by definition
> (1 + x) (1 + kx) by hypothesis and 2.3(4) (6 version)
= 1 + (k + 1) x + kx2 by A1–A4
> 1 + (k + 1) x by 2.3.
Hence the result follows by induction.
Proof. (1) and (2) are immediate from the definition and the fact that a > 0 iff −a < 0.
We can prove (3) and also (4), by using P3 to enumerate cases; recall too that (−a)(−a) =
a2 for any a.
For (5), note that (
−|a| 6 0 6 a = |a| if a > 0,
−|a| = a < 0 6 |a| if a < 0.
Now consider (6). Assume first that |a| 6 c, Then, by (5) and transitivity of 6, we get
−c 6 a 6 c. Conversely, assume −c 6 a 6 c. Then −a 6 c and a 6 c. Since |a| equals
either a or −a, we obtain |a| 6 c.
The case with < in place of 6 is handled similarly.
2.6. The Triangle Law and the Reverse Triangle Law. [BS 2.2.3 and 2.2.4] Thw Triangle
Law is also known as the Triangle Inequality.
(1) Let a, b ∈ R. Then
|a + b| 6 |a| + |b|.
(2) Let a, b ∈ R. Then
|a + b| > ||a| − |b|| .
12
so |a|−|b| 6 |a+b|, and likewise, reversing the roles of a and b, we get |b|−|a| 6 |b+a| = |a+b|.
Now use the fact that |c| is either c or −c always, and apply this with c = |a| − |b|.
Properties of the complex numbers are covered in the course Introduction to Complex
Numbers and not in Analysis I. But we shall deal with complex numbers occasionally, and it
is useful to record what does, and does not, hold in relation to arithmetic and inequalities.
We have noted earlier that
• (C; +, ·) is a field.
Here addition and multiplication are defined in the usual way, in terms of real and imaginary
parts, or in the case of multiplication, alternatively via polar representation. The axioms
A1–A4 follow from the corresponding axioms for R. The multiplication axioms M1–M4
are most easily verified using polar coordinates. Axiom Z holds since 0, 1 are real. Axiom
D holds by a straightforward, but tedious, calculation.
We highlight what does not transfer from R to C. The key point to note is that, unlike
R, the complex numbers do not carry a total order relation: we cannot define < on C in
a way which is compatible with arithmetic and such that, for any w, z ∈ C, exactly one of
w < z, w = z or z < w holds. Exercise: Prove this by considering i · i. Therefore
inequalities are off limits unless the quantities being compared are real.
Thus, for z = 3 + 4i√and w = 4 + i we may correctly say that Re z < Re w, that Im z > Im w,
and that |z| = 5 > 17 = |w|.
Both the Triangle Law and the Reverse Triangle Law extend to complex numbers (note
that the modulus of a complex number is real): for z, w ∈ C,
This section focuses on the order structure of R and exploits its interaction with the
arithmetic structure. In 4.3 we introduce the last of our axioms for R, the Completeness
Axiom. Throughout this section, and beyond, we shall make use of the axioms and results
in Sections 1 and 2, without spelling out the details.
On the basis of our axioms for arithmetic and order, we can’t distinguish between Q and
R. But you would claim they are different:
• in Q there is not a square root for 2 (standard proof, √
not included in lecture); but
• (you almost certainly believe that) there is a number 2 in R.
It will follow from the Completeness Axiom that 2 does have a square root in R (Theorem
4.10). Hence this axiom does distinguish R from Q (see 4.11).
Before presenting the Completeness Axiom we need some order-theoretic preliminaries.
Examples:
(a) R is not bounded above.
(b) Every element of R is an upper bound for the empty set.
(c) { s ∈ R | −4 6 s < 3 }: the numbers 3, 11, 1037 are upper bounds; 2.999999 is not an
upper bound. The set of lower bounds is (−∞, −4].
(d) { 1/n | n = 1, 2, . . . }: b is an upper bound for S iff b > 1 and b is a lower bound iff
b 6 0 (but can you justify this?).
(e) {1}: any b > 1 is an upper bound and any b 6 1 is a lower bound.
(f) Q: no upper bounds and no lower bounds.
(g) Let S := { x ∈ Q | x2 < 2 }. Then certainly S 6= ∅ and, for example, 3 is an upper
bound of S (why?).
Note that we can’t yet in every case here convincingly identify all the bounds.
Note the necessity for the exclusions here. The empty set has no supremum because it
has no least upper bound ((sup2) fails). A set which is not bounded above cannot have a
supremum because it has no upper bound ((sup1) must fail).
Example (sup and max compared): Let ∅ 6= S ⊆ R be such that sup S exists. Then S has
a maximum iff sup S ∈ S and then sup S = max S.
Similarly we say a non-empty set S which is bounded below has a minimum, min S, if
there exists s0 ∈ S such that s0 6 s for all s ∈ S.
15
Proof. 2.5(6) gives |x − a| < b iff −b < (x − a) < b and this holds iff a − b < x < a + b.
So by considering whether this holds for different values of b we can assess how good an
approximation x is to a.
4.11. Incompleteness of Q. The set Q of rationals does not satisfy the Completeness
Axiom with respect to the order it inherits from R. If it did,
T := { q ∈ Q | q > 0 and q 2 < 2 }
would have a supremum in Q. The proof in 4.10 works just as well for T as it does for S.
But we know there is no rational square root of 2.
4.12. Theorem (existence of nth roots). Any positive real number has a real nth root,
for any n = 2, 3, 4, . . ..
[Case of cube root of 2 is an exercise on Problem sheet 2.]
Proof. i) We know that inf S exists (by applying the completeness axiom to {−s : s ∈ S} as
in 4.6(b)). So by the approximation property with ε = 1 there is some n ∈ S such that
inf S 6 n < inf S + 1.
It is enough to show that inf S = n, since then inf S ∈ S and so inf S = min S. Assume for
a contradiction that n 6= inf S, so that n > inf S and hence n = inf S + ε where 0 < ε < 1.
By the approximation property again, there exists some m ∈ S such that
inf S 6 m < inf S + ε = n.
Since n > m we have n − m > 0 and so n − m > 1 because n − m is an integer1, so
n > inf S + 1, which contradicts our first inequality for n.
The proof of (ii) is similar.
1Strictly
speaking we have not proved this property of the integers. It can be proved by observing that a
strictly positive integer is a natural number and using induction to deduce that it must be at least 1.
17
5. An aside: countability
The density properties recorded in 4.15 are important, the more so because, as we shall see
shortly, R is ‘a bigger set’ than Q: the set Q is countable whereas the set R is uncountable.
The objective in Analysis I is not to present a crash course in set theory but to give the
minimum amount of information on countability necessary to distinguish between countable
and uncountable sets in the context of the real numbers. See the supplementary notes on
Countability for an informal account of this topic which goes beyond the Analysis I syllabus.
These notes may be of interest to those who want to go deeper than Section 5 does and in
particular to Maths/Phil students.
We call a set A
• countably infinite if A ≈ N;
• countable if A 4 N;
• uncountable if A is not countable.
[Warning: Some authors use ‘countable’ to mean what we call ‘countably infinite’.]
FACTS: (see supplementary material on countability)
(1) A is countable (that is, A 4 N) iff A is finite or countably infinite.
(2) If A 4 B and B 4 A then A ≈ B.
Proof. (a) is immediate. For (b): the successor function, n 7→ n + 1, is a bijection from N
to N \ {0}. To prove (c), note that the map 2k + 1 7→ k is injective and maps the given set
onto N. Now consider (d). We can define a bijection f from Z to N by
(
−2k if k 6 0,
f (k) =
2k − 1 if k > 0.
Proof. Define (
2f (x) if x ∈ A,
h(x) =
2g(x) + 1 if x ∈ B.
Then h is an injection from A ∪ B to N.
Claim 2: A × B is countable.
19
Proof. Define h : A × B → N by
h((a, b)) = 2f (a)+1 3g(b)+1 for a ∈ A, b ∈ B.
Then uniqueness of factorisation in N>0 implies h is injective.
We want our treatment of sequences, and a bit later, of series not to be divorced from
the functions you were introduced to at school and regularly encounter in other courses:
trigonometric functions, exponential functions, logarithms, and general powers. Accordingly
we’d like to involve such functions in examples and exercises. So, for now, we shall take
the existence and the properties of these functions for granted, and use them freely. Later
you will see formal definitions of these various functions and rigorous derivations of their
(familiar) properties.
When we use logarithms these will always be to base e. We adopt the notation log x for
loge x, rather than ln x.
Recall that, for a > 0 and x ∈ R, one defines ax = ex log a .
6. Sequences
This section covers the rudiments of the theory of convergence of sequences. An ample
supply of worked examples is included in these webnotes. Examples omitted from lectures
are recommended for self-study.
20
6.3. Manufacturing new sequences. We can form new sequences from given ones ‘a term
at a time’: given sequences (an ) and (bn ), we have sequences (an + bn ), (−an ), (an bn ) and,
provided every bn is non-zero, (an /bn ). Also we can form (can ), for any constant c, and
(|an |).
Examples: let an = (−1)n and bn = 1 for all n. Then
The key notion in this section is that of convergence of a sequence (an ). We want to
analyse how the terms an of the sequence behave as n gets ‘arbitrarily large’ and specifically
whether or not the terms approach ‘arbitrarily closely’ some ‘limiting value’ L. For this we
need to convert these informal ideas into precise, formal, ones.
6.4. Tails. As regards the long-run behaviour of the terms of a sequence (an ) as n gets
arbitrarily large we don’t care what the values of the first few terms are, or what the first
10 million terms are, . . . . The notion of a tail of (an ) will allow us to capture this idea.
Given a sequence (an ) and any k ∈ N we can form a new sequence (bn ) by chopping off
the first k terms a1 , . . . , ak of (an ) and relabelling. That is, bn = an+k for all n. We call (bn )
a tail of (an ).
6.5. Capturing ‘arbitrarily close to’ via ε. From 2.5(6) (see also 4.8), we have, for
x, a ∈ R and b > 0,
|x − a| < b ⇐⇒ a − b < x < a + b
and this condition captures the statement that x lies within a distance b of a. In formulating
notions of limits and convergence we are interested in allowing b to be very small—arbitrarily
small—and shall use the customary symbol ε to denote a strictly positive real number playing
this role.
6.6. Convergence of a real sequence. Let (an ) be a sequence of real numbers and let
L ∈ R. Then we say that (an ) converges to L (notation: an → L (as n → ∞)) if
Here N can, and almost always will, depend on ε. Note that we can replace ‘n > N ’ by
n > N and/or ‘|an − L| < ε’ by |an − L| 6 ε in this definition without changing the meaning
(WHY?). However it is crucial that ε should be strictly greater than 0.
When an → L we say that L is the limit of (an ) [in 6.13 we’ll show it must be unique]
and we write
L = lim an or just L = lim an .
n→∞
Proof. (i) Let k ∈ N and let bn = an+k for n > 1. Assume (an ) converges to L. Then
∀ε > 0 ∃N ∀n > N |an − L| < ε.
This implies |bn − L| = |an+k − L| < ε for all n > N because then n + k > n > N .
(ii) Since (bn ) converges there exists L such that
∀ε > 0 ∃N 0 ∀p > N 0 |bp − L| < ε.
That is, p > N implies |ap+k − L| < ε. Define N = N 0 + k. Then n > N implies that
0
6.8. Examples: convergence established directly (but messily) from the defini-
tion.
(a) The basic fact that
1/n → 0 as n → ∞
is simply the Archimedean Property in new clothes. To see this note that the Archimedean
Property is the statement that, given ε > 0, there exists N ∈ N such that 1/N < ε.
Then for n > N we have |1/n − 0| = 1/n 6 1/N < ε.
See 6.20 for a discussion of the limiting behaviour of arbitrary powers of n.
1
(b) Let an = 1 + (−1)n √ . We claim an → L where L = 1. Let ε > 0. Then
n
n 1 1
|an − 1| = (−1) √ < ε ⇐= √ < ε
n n
1
⇐= n > N, where we choose N ∈ N with N > .
ε2
Here , and likewise below, the first ⇐= may be read as ‘if’ and any subsequent ones as
‘and hence if’.
1
(c) Let an = . Let ε > 0. Then
n2−n+1
1
|an − 0| = 2 < ε ⇐= n2 − n + 1 > ε−1
n −n+1
⇐= (n − 21 )2 > ε−1 − 3
4
q
1
⇐= n > N, where N ∈ N and N > 2
+ ε−1 − 34 ,
where we may assume without loss of generality (see 6.9(4)) that ε < 4/3. So an → 0.
n sin n2
(d) Let an = 3 . Let ε > 0. Then
3n − n − 1
n| sin n2 | n
|an − 0| = 3 < ε ⇐= 3 <ε since | sin x| 6 1 for all x
3n − n − 1 3n − n − 1
⇐= n/n3 < ε since n3 > n and n3 > 1
√
⇐= n > N where N > 1/ ε.
So an → 0. Note the need for care in handling inequalities in this example.
23
6.9. Testing a sequence for convergence: remarks and technical tips. Consider the
situation in which we have a sequence (an ) and a candidate limit L, and we wish to show
that an → L, that is, that it satisfies the ε-N condition in 6.6.
(1) We require N such that
n > N =⇒ |an − L| < ε.
We do NOT need
n > N ⇐⇒ |an − L| < ε.
This means we do not need to find the smallest N possible when establishing conver-
gence. Any N that works will do. This allows us in many cases to simplify calculations
by replacing complicated expressions by simpler ones before trying to write down a
suitable N .
(2) Beginners often give back-to-front arguments when seeking to prove that a sequence
(an ) tends to some limit L. Note the direction of the implication signs in Examples
6.8. Reading downards, we are working towards finding a suitable N . Once such an N
has been identified, the argument can be re-presented, going from bottom to top and
using forward implication
√ signs. Let’s carry this out for the sequence (an ) in 6.8(b),
where an = ((−1)n / n), treating the original presentation as rough work. Take ε > 0.
Choose N ∈ N with N > 1/ε2 . Then
1
n > N =⇒ n > 2
ε
1
=⇒ < ε2
n
n 1
=⇒ (−1) √ < ε
n
=⇒ |an − 1| < ε.
In the examples in 6.8 most instances of ⇐= could be replaced by ⇐⇒, but not all
can be (see (d)); (1) says that none needs to be.
(3) We have asked that N ∈ N. But it’s good enough to find X ∈ R such that n > X
implies |an − L| < ε. If X exists then we can choose N > X and N ∈ N (since N is not
bounded above, as we proved in 4.13(i)).
(4) The smaller ε is, the greater challenge we have to find a corresponding N in general.
Turning this around, we see that in establishing convergence we may without loss of
generality restrict to values of ε such that ε < 1 (or ε < η, where η is some fixed positive
number). We did in this in 6.8(c)). It’s small values of ε that matter.
(5) Facility with inequalities is a valuable skill!
It is already clear from Examples 6.8(c),(d) that finding an explicit N (or X) for a given
ε > 0 can be tiresome and messy. The following result is elementary. It is useful in two
ways: it allows us
• to simplify ε-N proofs;
• to take advantage of known limits to find the limiting values of other sequences.
6.10. Sandwiching Lemma (simple form). Let (bn ) and (cn ) be real sequences. Assume
cn → 0 and that 0 6 bn 6 cn for all n. Then bn → 0.
Proof. Let ε > 0 and pick N so that |cn − 0| < ε for all n > N . Then, for n > N ,
−ε < 0 6 bn 6 cn = |cn | < ε.
24
[What this says is that, given ε, an N that works for (cn ) also works for (bn ).]
(b) 0 6 2−n 6 1/n for all n (by induction) and 1/n → 0. Hence (2−n ) converges to 0 by
the Sandwiching Lemma.
n
(c) Let an = √ . Then an → 1. To prove this, note that
n2 + 1
√ √
n2 + 1 − n n2 + 2n + 1 − n (n + 1) − n 1 1
0 6 |an − 1| = √ 6 √ = √ =√ < .
n2 + 1 n2 + 1 n2 + 1 n2 + 1 n
Now apply the Sandwiching Lemma, once again using the fact that 1/n → 0.
(d) [Example 6.8(d) revisited] The proof given earlier is a sandwiching argument from
scratch. The Sandwiching Lemma can be applied with bn = |an | and cn = 1/n.
6.12. More examples (two important limits, employing some useful techniques).
(a) Let |c| < 1, where c is constant. We claim cn → 0. We can write |c| = 1/(1 + y) where
y > 0. Take ε > 0. By Bernoulli’s inequality,
1 1 1
|c|n = n
6 < < ε if n > N,
(1 + y) 1 + ny ny
where N ∈ N is chosen such that N > 1/(yε).
n
(b) Let an = . We surmise that an → 0. Let ε > 0. Now
2n
n n n
|an − 0| = n = n
= n
(by the binomial theorem)
2 (1 + 1) 1 + n + 2 + ··· + 1
2n n
6 (if n > 2, by retaining the 2
term)
n(n − 1)
<ε provided n − 1 > 2/ε.
So we choose N ∈ N so that N > 1 + 2/ε.
Now we give a significant theoretical result. The proof gives an illustration of working
with the ε-N definition of convergence.
6.13. Theorem (uniqueness of limits). Let (an ) be a sequence and suppose that an → L1
and an → L2 as n → ∞. Then L1 = L2 .
Proof. Suppose L1 6= L2 . Take ε := |L1 − L2 |. Then ε > 0. So ε/2 > 0 and hence there exist
N1 and N2 such that
n > N1 =⇒ |an − L1 | < ε/2,
n > N2 =⇒ |an − L2 | < ε/2.
25
The next group of results concerns the interaction of limits with modulus and inequalities.
6.14. Proposition (limits and modulus). Assume that (an ) is a sequence which converges
to L. Then (|an |) converges too, to |L|.
6.15. Limits and inequalities. Let (an ) and (bn ) be real sequences.
Preservation of weak inequalities: Assume that an → L and bn → M and that an 6 bn
for all n. Then L 6 M .
6.16. A general Sandwiching Lemma. Assume that (xn ), (yn ) and (an ) are real sequences
such that xn 6 an 6 yn for all n. Assume that lim xn = lim yn = L. Then (an ) converges
to L.
Proof. (Outline) Given ε > 0 we can find N such that for all n > N we have |xn − L| < ε
and |yn − L| < ε. Then, for n > N ,
L − ε < xn 6 an 6 yn < L + ε,
so |an − L| < ε.
The next result shows us that any convergent sequence has a special property, that of
being bounded. The result will be useful in some technical proofs later, and also, in its
contrapositive form, provides a way to show that certain sequences fail to converge.
26
n2
(b) Let an = (−1)n . Then (an ) diverges.
n2 + 1
The result is plausible, but this is an awkward example to handle slickly from first
principles. We’ll give a proof later, once we have developed an efficient method.
6.19. Infinity. Let an be a sequence of real numbers. We say ‘an tends to infinity’ and write
an → ∞ as n → ∞ if
∀M ∈ R ∃N ∈ N ∀n > N an > M.
Similarly we write bn → −∞ if
∀M ∈ R ∃N ∈ N ∀n > N bn < M.
(Here we tend to think of M as being a very large positive/negative number.)
Here the symbol ∞ provides a convenient notational shorthand. Do NOT treat ∞ as
though it were a real number. Remember that in our convergence definition in 6.6 we
demanded that the limit L belong to R. See further discussion of infinite limits in Section 7.
Examples
(a) Let an = n2 − 1000. Then an → ∞. To prove >0
√ this let M ∈ R . Then an > M for all
( choose N ∈ N such that N > M + 1000—possible by 4.13(i).
n > N if we
n if n is odd,
(b) Let an =
0 if n is even.
27
If it were true that an → ∞, then (taking M = 1) we could find N such that n > N
implies an > 1. But a2N = 0 and we have a contradiction.
6.21. Complex sequences. As noted already, the definition of convergence and much of
the theory of convergence of real sequences carry over in the obvious way to sequences of
complex numbers. In particular we say a sequence (zn ) converges to L (where now L ∈ C) if
∀ε > 0 ∃N ∈ N ∀n > N |zn − L| < ε.
Moreover, the limit is unique if it exists. Uniqueness is proved as in 6.13.
Recall the remarks in Section 3 concerning complex numbers.
• Inequalities: don’t try to write w < z when w, z ∈ C, not both real !
• The Triangle and Reverse Triangle Laws do hold for complex numbers.
• Simple sandwiching is valid in the following form: suppose (wn ) and (zn ) are complex
sequences such that |wn | 6 |zn | and zn → 0, then wn → 0. [Note the moduli ! ]
Proof. The proof of uniqueness is exactly the same as in the real case.
Now consider the first assertion. For =⇒ use the fact that |xn | 6 |zn | and |yn | 6 |zn |
and sandwiching. The proof of ⇐ is a definition-chase and left as an exercise (ideas from
8.3 are useful).
28
7. Subsequences
This short section introduces the notion of a subsequence of a (real or complex) sequence.
The results we obtain here usefully enlarge our armoury of techniques for establishing con-
vergence/divergence. Deeper results involving subsequences are given in Section 10.
7.1. Subsequences. Let (an )n>1 be a (real or complex) sequence. Informally, a subsequence
of (an )n>1 is a sequence (br )r>1 whose terms are obtained by taking infinitely many terms
from (an )n>1 , in order. So, for example, if (an ) = (1, 2, 3, 4, . . .) then:
(2, 4, 6, . . .) is a subsequence of (an ) (it is the sequence (a2n )),
(2, 22 , 24 , . . .) is a subsequence of (an ) (it is the sequence (a2n )),
(6, 4 . . .) is not a subsequence of (an ) (terms not in correct order),
(2, 4, 0, 0, . . .) is not a subsequence of (an ) (not all terms are terms of (an )),
(1, 2, 3, 4, . . . , 2015) is not a subsequence of (an ) (not a sequence).
Formally, a subsequence (br )r>1 of the sequence (an )n>1 is defined by a map f : N → N
such that f is strictly increasing (meaning that r < s implies f (r) < f (s)), so that
br := anr , where nr = f (r).
Expressing this another way, we have a infinite sequence of natural numbers
n1 < n2 < n3 < . . .
29
Proof. (i) Take ε > 0. Choose N ∈ N such that n > N implies |an − L| < ε for all n > N .
In particular, nr > N implies |anr − L| < ε. Since r 6 nr this holds whenever r > N . Thus
we have proved that
∀ε > 0 ∃N ∈ N ∀r > N |anr − L| < ε.
8.2. Facilitating the construction of ε-N proofs: making proofs less fiddly.
(1) As noted earlier, in proving convergence it is sufficient to consider, for example, ε < 1;
it is only small values of ε that we need to consider.
(2) (an ) converges to L if, for some constant K > 0,
∀η > 0 ∃N ∈ N ∀n > N |an − L| < Kη.
Here it is crucially important that K is constant—it must not involve n.
Proof. Take ε > 0. Apply the given condition with η := ε/K to get the standard
condition for convergence.
(3) Remember that (an ) converges provided it has a convergent tail; recall the Tails Lemma,
6.7. [We needn’t even fuss if we have a sequence whose first few terms are not defined.]
We’ll split the AOL results into two groups. Here are the ones which are easier to prove.
8.6. Theorem: (AOL), Part II. Assume that (an ) and (bn ) are (real or complex) se-
quences and assume that (an ) converges to L and (bn ) converges to M . Then the following
hold as n → ∞.
(vi) (product) an bn → LM .
(vii) (reciprocal) If M 6= 0, then 1/bn → 1/M .
(viii) (quotient) an /bn → L/M if M 6= 0.
[In (vii) and (viii), we may need to restrict to a tail of, respectively, (1/bn ) and (an /bn ) to
get well-defined terms. This will always be possible by 8.5(2).]
Proof. Write p(n)/q(n) in the form nk−` [· · · ]. The quantity in square brackets tends to
1, by the (AOL) results for constants, scalar multiplication, product and quotient. If
k 6 ` the required result then follows by (AOL) for product. For the case k > `, note
q(n)/p(n) → 0, by interchanging the roles of p and q above. Then use Proposition 8.8,
noting that q(n)/p(n) is positive for large n.
(b) [A useful limit] Let a > 1 and m ∈ N>0 . Then nm /an → 0 as n → ∞. [The sequence
(n2−n )n>1 considered in 6.7(b) is a special case.]
Proof. Write a = 1 + b, where b > 0. Then, by the binomial theorem,
nm nm
= .
1 + · · · + nk bk + · · · + bn
(1 + b)n
All the terms in the denominator are positive so if we drop all but one of these we make
the fraction bigger. We elect to retain the term with k = m + 1, where we assume
n > m + 1. We have
n n! n(n − 1) · · · (n − k + 1)
= = .
k (n − k)!k! k!
Then
nm nm
(m + 1)!
06 6 .
(1 + b)n bm+1 n(n − 1) · · · (n − m)
The expression on the right-hand side tends to 0: the first term is a constant, and the
second tends to 0 by (a). Now use simple sandwiching.
33
8.10. A ‘true-or-false?’ worked example. Let (an ) be a real sequence such that an → ∞
as n → ∞. To decide, with a proof or counterexample as appropriate, whether an /bn → ∞
under each of the following assumptions.
(a) Assume that (bn ) is a bounded sequence with bn 6= 0. Consider an = n and bn = (−1)n .
Then an /bn 6→ ∞.
(b) Assume that (bn ) is a bounded sequence with bn > 0 for all n (or for n > k for some
k). Then an /bn → ∞ is true. There exists K (constant) such that if 0 < bn 6 K, and
so also 1/bn > 1/K, for a tail of (bn ). Since an → ∞, given M , we have an > M K for
large n and hence an /bn > (M K)/K = M .
(c) Assume that (bn ) is a sequence which converges to L > 0. Then an /bn → ∞. Note that
(bn ) is bounded because it is convergent and that bn > 0 for large n by 8.5(2). Now
appeal to (b).
Don’t expect AOL results to work when infinite limits come into play. Note also Problem
sheet 3, Question 5.
8.11. Orders of magnitude. When looking for a candidate limit for a given sequence (an ),
an obvious strategy is to look for the dominate components in an , which can be expected
to dictate the sequence’s behaviour in the long term. Usually for this we need to appreciate
the relative magnitudes of the terms as n becomes large.
Examples
n3 − 107 n
(a) Let an = . For large n, the dominant term in the numerator of an is n3
n5 + 6n + 1
and that in the denominator is n5 . So we expect the sequence to behave like (n−2 ), and
to have limit 0. Indeed, this is what exactly what an argument using the Algebra of
Limits formalises.
√
(c) Let an = sin(nn )/ n. The oscillatory sine function does not assist in getting conver-
n
gence of an to 0 but because | sin x| 6 1 for all x, the
√ rapid growth of n is an irrelevance.
To get the limit, just use sandwiching: |an | 6 1/ n → 0 as n → ∞.
As a rule of thumb, when it comes to the behaviour of functions f (x) for large x:
8.12. The O and o notation. Let (an ) and (bn ) be real or complex sequences. We write
an = O(bn ) as n → ∞ if there exists c such that for some N
n > N =⇒ |an | 6 c|bn |.
34
Examples
(a) n + 106 = O (n2 ) is true but n2 = O(n) is false;
(b) n = o (n2 ) is true;
(c) sin n = O (1) is true;
(d) by 8.9(b) when a > 1 and m ∈ N we have nm = o (an ) as n → ∞.
The symbol ≈, used to indicate ‘approximately the same size as’, is not precise (how close
an approximation is intended?) and so ≈ is outlawed in this course. O, by contrast, has a
precise meaning.
The O notation is useful in particular for simplifying calculations while maintaining rigour.
Given a complicated an , consideration of dominant terms in the expression defining an ) can
help us a simpler expression bn with an = O(bn ).
9. Monotonic Sequences
So far, in order to prove a sequence converges, we have had to identify a candidate limit
at the outset. There is a very important class of real sequences which can be guaranteed to
converge to a limit L ∈ R, or to tend to ∞ or to −∞.
For the converse we use the fact that any convergent sequence is bounded (6.17).
For (ii), note that (an ) is monotonic decreasing and bounded below iff (−an ) is monotonic
increasing and bounded above.
35
Points to note
• A real sequence which has a tail which is monotonic increasing converges iff it is bounded
above.
• A real sequence which is monotonic increasing and not bounded above tends to ∞.
Proof. Given M there exists N such that aN > M . But then n > N implies an > M ,
since (an ) is monotonic increasing.
(4) Putting the pieces together: The sequence (an ) is bounded below by 0 and
(an )n>2 is monotonic decreasing by (3). By the Monotonic
√ Sequence Theorem,
an → L for some (finite) limit L. By (1), L must be c.
√
This example proves the √existence of c. Where this argument differs from the fiddly
one used earlier to show 2 exists is in the availability now of the Algebra of Limits.
Observe that both the new proof and the old one use the Completeness Axiom.
In this section we obtain important general results about sequences which are not mono-
tonic. Since monotonic real sequences behave so well, the following theorem provides welcome
information.
37
10.1. The Scenic Viewpoint Theorem. Let (an ) be a real sequence. Then (an ) has a
monotonic subsequence.
Proof. We consider the set V = { k ∈ N | m > k =⇒ am < ak }. This is the set of “scenic
viewpoints” (also known as “peaks”)—given k ∈ V , looking towards infinity from a point at
height ak , no higher point would impede our view.
Case 1: V is infinite. Then the elements of V can be enumerated as k1 < k2 < . . . .
Then (akr ) is a subsequence of (an ) and
r > s =⇒ kr > ks =⇒ akr < aks
that is, (akr ) is monotone decreasing.
Case 2: V is finite. Let N be such that every element of V is < N . Then m1 = N
is such that m1 ∈/ V , so there exists m2 > m1 with am2 > am1 . Since m2 ∈
/ V , there
exists m3 > m2 such that am3 > am2 . Proceeding in this way we can (inductively)
generate a monotonic increasing sequence (amk ).
10.2. Bolzano–Weierstrass Theorem (for real sequences). Let (an ) be a bounded real
sequence. Then (an ) has a convergent subsequence.
Proof. By the Scenic Viewpoint Theorem 10.1, (an ) has a monotonic subsequence which is
also bounded. By the Monotonic Sequence Theorem, this subsequence converges.
Neither the Monotonic Sequence Theorem nor the Scenic Viewpoint Theorem extends to
complex sequences, but the Bolzano–Weierstrass Theorem does.
Proof. Write zn = xn + iyn . Let M be such that |zn | 6 M for all n > 1. Then by the
definition of modulus in C, we have |xn | 6 M and |yn | 6 M , for all n > 1. Therefore by
the real BW Theorem, (xn )n>1 has a convergent subsequence, say (xnr )r>1 . Let wr := ynr ,
for r > 1. Then the subsequence (wr )r>1 , being a subsequence of the bounded sequence
(yn )n>1 , is itself bounded. Choose a subsequence (wrs )s>1 of (wr )r>1 which is convergent.
Then (ynrs )s>1 converges and so does (xnrs )s>1 , since it is a subsequence of a convergent
sequence. Finally, (znrs )s>1 is convergent since the sequences of real and imaginary parts
converge.
10.4. Cauchy sequences. The idea behind Cauchy sequences is that if the terms of a
sequence (an ) are ultimately arbitrarily close to one another, then there should be a value
L to which they must converge. This would be valuable to know, since up till now we have
needed to identify a candidate limit in advance (except for bounded monotonic sequences,
where we may be able to find the limit from a given, or constructed, recurrence relation).
Definition. Let (an ) be a real or complex sequence. Then (an ) is a Cauchy sequence if
it satisfies the Cauchy condition:
∀ε > 0 ∃N ∈ N ∀m, n > N |an − am | < ε.
It is not sufficient in the Cauchy condition just to consider adjacent terms, that is, only
to consider m = n + 1.
38
10.6. Theorem (Cauchy Convergence Criterion). Let (an ) be a (real or complex) se-
quence. Then
(an ) is convergent ⇐⇒ (an ) is a Cauchy sequence.
Proof. =⇒ : By 10.5(2).
⇐=: By 10.5(1) and the BW Theorem, 10.2 or 10.3, (an ) has a convergent subsequence.
Now appeal to 10.5(3).
Note on notation: We have two sequences in play at the same time here: the sequence
(ak ) of the terms
P of the series and the sequence (sn ) of partial sums. To avoid writing
(erroneously!) nn=1 an , we have introduced a new (dummy) variable k to label the terms.
1 P
11.5. Example: the harmonic series, . Let
k>1
k
1 1 1
sn = 1 + + + · · · + (n > 1).
2 3 n
We claim (sn ) is not a Cauchy sequence, so (sn ) does not converge and therefore
X1
diverges.
k>1
k
Proof. Consider
1 1
|s2n+1 − s2n | = + · · · + n+1
+12n 2
1
> n+1 · (2n+1 − 2n ) ((smallest term) × (number of terms))
2
1
= .
2
Therefore (sn ) is not Cauchy, so fails to converge (by 10.6)).
P P 1
This example shows also that ak → 0 does not imply k>1 ak converges: k>1 provides
k
a counterexample.
Later (12.13(c)) we’ll see that sn grows about as fast as log n, in that sn − log n tends to
a finite constant.
A particularly amenable class of series will be those whose partial sum sequences are
monotonic, thanks to the Monotonic Sequence Theorem.
41
Proof. By the fact that sn − sn−1 = an (n > 2) and Monotonic Sequence Theorem, 9.2.
Proof. Let
sn = a1 + · · · + an and tn = b1 + · · · + bn .
Then sn 6 Ctn for all n. Since (tn ) converges, it is bounded above, by T say. Hence sn 6 T
for all n, and Theorem 11.6 applies.
(a) Take ak = 1/k 2 and bk = 1/(k(k + 1)). Then 0 6 ak 6 2bk and we proved earlier that
P P 1
k>1 bk converges. Hence k>1 2 converges.
k
P
(b) Let ak = 1/k! and bk = 1/(k(k P + 1)). Then 0 6 ak 6 bk and hence the series k>1 1/k!
converges by comparison with P k>1 1/(k(k + 1)). Hence the series defining e converges
(note this series is not exactly k>1 ak because the latter has one fewer term at the
front but this does not affect convergence).
11.9. Theorem (Cauchy criterion for convergence of P a series). Let (ak ) be a real or
complex sequence with partial sum sequence (sn ). Then k>1 ak converges if and only if
11.10.
P Absolute convergence. Let (aP k ) be a sequence of real or complex numbers. We
say k>1 ak converges absolutely if k>1 |ak | converges.
P
Note that k>1 |ak | is a series of non-negative terms, to which Theorem 11.6 applies. Thus
the following theorem is very useful.
42
−p
P
(a) k>1 k diverges if p 6 1;
Proof. For p 6 0, the terms do not tend to 0. For p = 1 we have the harmonic series.
For 0 < p < 1, use the contrapositive of the simple Comparison Test, with ak = k −1
and b = k −p .
P k −p
(b) k>1 k converges if p > 2.
Proof. Use simple Comparison Test, comparing with k>1 1/k 2 , which we proved ear-
P
lier is convergent.P
(c) [Looking ahead] k>1 k −p converges if 1 < p < 2.
Proof deferred until we have available the Integral Test, which allows us to handle
all values of p > 0 in a uniform way; see 12.13(a).
In summary:
In this section we shall obtain some very useful tests for convergence/divergence of series.
For testing series of non-negative terms and so for testing for absolute convergence
(from which convergence follows):
• Comparison Test (limit form);
• D’Alembert’s Ratio Test;
• Integral Test.
For testing series (−1)k−1 uk , where uk > 0:
P
We shall present the Alternating Series Test first. It is quite special and there are close
connections with material in the previous section.
(−1)k−1 uk converges if
P
12.1. Leibniz’ Alternating Series Test. The series
(i) uk > 0,
(ii) uk+1 6 uk ,
(iii) uk → 0 as k → ∞.
k−1 1
P
(a) By AST, k>1 (−1) converges.
k
Here
Pwe have an example of a series which converges but fails to converge absolutely,
since k>1 1/k diverges (recall 11.5).
P k−1 √1
(b) By AST, k>1 (−1) converges. This is a useful series for counterexamples.
k
12.4. Comparison Test, limit form. Let ak and bk be strictly positive and assume that
ak
→ L, where 0 < L < ∞.
bk
Then
P P
ak converges ⇐⇒ bk converges.
Proof. In the limit definition take ε = L/2 and choose N such that, for k > N ,
ak
− L < 1 L and hence 1 L < ak < 3 L.
bk 2 2 bk 2
P P
Then, restricting to tails, bk convergent implies that aPk is convergent, by the simple
P
Comparison Test with C = 3L/2. In the other direction, ak convergent implies bk
−1
converges by comparison, because 0 < bk < 2L ak for large k.
Note: It is crucial that the terms ak and bk are (ultimately) strictly positive and that L is
non-zero and finite.
The Comparison Test in either form is not ‘internal’, in that we have to produce a suitable
series with which to compare. We next obtain some important tests which don’t have this
disadvantage.
12.6. D’Alembert’s Ratio Test, for series of strictly positive terms. Let ak > 0.
Assume that
ak+1
lim exists and equals L.
k→∞ ak
Then
P
06L<1 =⇒ ak converges;
P
L>1 =⇒ ak diverges;
L=1 =⇒ the test gives no result.
(Here it is permissible, and useful, to allow L = ∞ as a possible limit and to treat it as > 1
for the purposes of stating the test.)
• For a series with ‘gaps’, leave out the zero terms and relabel before applying the test.
For example, consider
1
if k is of the form 2m for some m > 1,
ak = 2k!
0 otherwise,
P
so ak (starting from k = 1) looks like
1 1 1
0 + 1! + 0 + 4! + 0 + 0 + 0 + 8! + 0 + . . . .
2 2 2
m
We cannot apply the Ratio Test directly to k>1 ak but we can apply it to m > 11/2(2 )! .
P P
12.9. Testing for absolute convergence: a corollary to the Ratio Test 12.6. Let ak
be non-zero real or complex numbers and assume that
ak+1
lim exists and equals L.
k→∞ ak
The (corollary to the) Ratio Test is particularly useful for testing series
X X
ck xk (x real) and ck z k (z complex)
for convergence. Here we regard x or z as a variable, so that the series, provided it is
convergent, will define a function. The whole of Section 13 is devoted to such power series
and their properties. There we give further examples of the use of the Ratio Test in important
particular cases.
12.10. The Integral Test: preamble. P Here we shall analyse the the behaviour of the
partial sum sequence (sn ) of a series f (k) by comparing it with the sequence (In ) where
Z n
In = f (x) dx,
1
where f : [1, ∞) → [0, ∞) is a suitable function. We shall need to make use of standard
properties of integrals (integration is treated in Analysis III in Trinity Term). We assume
the following. On a suitable class of integrable functions:
(a) integration preserves 6;
R k+1
(b) k c dx = c, for any constant c;
Rc Rb Rc
(c) intervals slot together: a = a + b for a < b < c.
47
12.11. Integral Test Theorem. Assume that f is a real-valued function defined on [1, ∞)
with the following properties:
(i) f is non-negative and decreasing;
R k+1
(ii) k f (x) dx exists for each k > 1 [for future reference: this holds if f is continuous].
Let
n
X Z n
sn = f (k) and In = f (x) dx.
k=1 1
Proof. Note that, because f is decreasing, f (k) > f (x) > f (k + 1) for all x ∈ [k, k + 1]. Now,
by properties (a) and (b) above,
Z k+1
f (k + 1) 6 f (x) dx 6 f (k) (k = 1, 2, . . .) .
k
So we have
Z 2
f (2) 6 f (x) dx 6 f (1)
1
Z 3
f (3) 6 f (x) dx 6 f (2)
2
··· ··· ···
Z n
f (n) 6 f (x) dx 6 f (n − 1).
n−1
Then
n
X Z n
0 6 f (n) 6 f (k) − f (x) dx 6 f (1),
k=1 1
Therefore (σn ) is monotonic decreasing and bounded below, and hence converges to a limit σ,
where 0 6 σ 6 f (1).
12.12. Corollary: the Integral Test. Assume that (as in 12.11), f : [1, ∞) → [0, ∞) is
R k+1 P
monotonic decreasing and such that k f (x) dx exists for each k. Then f (k) converges
if and only if (In ) converges.
Proof. This is immediate from elementary properties of limits and the fact that (sn − In )
converges.
48
12.13. Applications of the Integral Test and Integral Test Theorem. Here we assume
not just properties of integrals but also how to evaluate standard integrals.
P −p
(a) k (p > 0), definitively. Take f (x) = x−p . Then f is non-negative and decreasing
on [1, ∞). We have
1 1
−p+1 n
(n−p+1 − 1) if p 6= 1,
Z n
−p x 1
=
In = x dx = −p + 1 −p + 1
1 log n if p = 1.
P −p
Hence (In ) converges (to a finite limit) iff p > 1. Therefore, as claimed in 11.12, k
converges
P −p if p > 1 and diverges if 0 < p 6 1. [If p < 0 the Integral Test does not apply
but k diverges because k −p 9 0.]
(b) The Integral Test applied with f (x) = 1/(x log x) (note that the conditions for the test
are met, except that we need to start from x = 2 rather than x = 1) implies that
P
k>2 1/(k log k) diverges. This is, perhaps, a little surprising.
(c) Euler’s constant, γ. Apply the Integral Test Theorem in the special case that f (x) =
1/x; certainly f is non-negative and monotonic decreasing, and its integral exists over
any interval [k, k + 1]. We have that
1 1 1
γn := 1 + + + · · · + − log n → γ,
2 3 n
where γ is a constant between 0 and 1. This shows that the partial sums of the divergent
harmonic series tend to infinity slowly—about as fast as log n. The constant γ, known
as Euler’s constant, is rather mysterious: it remains unknown whether γ is rational or
irrational.
1 1 1
(d) A series for log 2. Let sn = 1 − + + · · · + (−1)n−1 . Then
2 3 n
1 1 1
s2n = 1 − + + · · · −
2 3 2n
1 1 1 1 1 1
= 1 + + + ··· + −2 + + ··· +
2 3 2n 2 4 2n
= (γ2n + log(2n)) − (γn + log n)
= log 2 + γ2n − γn
→ log 2.
12.14. Example: exploiting the existence of Euler’s constant. [details omitted from
lectures] Consider
1 1 1 1
1 − + − + − ... and
2 3 4 5
1 1 1 1 1 1 1
1+ + − + + + − + ...,
3 5 2 7 9 11 4
so the second series contains the same terms as the first, but in a different order. We analyse
the limiting behaviour of the second series. The terms come in groups of four, three positive
terms followed by one negative one, so we first look at the sum of 4n terms, in n groups:
1 1 1 1 1 1 1 1 1 1 1
s4n = 1 + + − + + + − + ··· + + + −
3 5 2 7 9 11 4 6n − 5 6n − 3 6n − 1 2n
1 1 1 1 1 1 1 1 1
= 1 + + + ··· + − + + ··· + − + + ··· +
2 3 6n 2 4 6n 2 4 2n
1 1
= (γ6n + log(6n)) − (γ3n + log(3n)) − (γn + log n)
2 2
49
36n2
1 1 1
= log + γ6n − γ3n − γ2n
2 3n2 2 2
1 1
→ s := log 12 = log 2 + log 3.
2 2
So far we have only looked at a particular subsequence of (sn ). Not good enough! But it is
easy to remedy the omission:
1
s4n+1 = s4n + → s,
6n + 7
1
s4n+2 = s4n+1 + → s,
6n + 9
1
s4n+3 = s4n+2 + → s.
6n + 11
We deduce that sn → s = log 2 + 21 log 3 (using an extension of the result of Problem sheet
4, Q.2(a)).
So, by changing the order P
of the terms we have changed the value of the sum! We showed
that the alternating series (−1)k−1 1/k has sum log 2, whereas the ‘3 pluses, 1 minus’
rearrangement has sum log 2 + 21 log 3.
: N>0 → N>0 be a
P
12.15. Definition: rearrangement.P Take any series ak and let gP
bijection. Let bk = ag(k) . Then bk is said to be a rearrangement of ak .
12.18. A ‘true-or-false?’ worked example. There are many conjectures one might make
about convergence of series in general. A few, some true, some false, can be decided even on
our limited treatment of series, and limited examples, thus far.
P P
(a) For ak , bk real, ak 6 bk and bk convergent implies ak convergent.
FALSE: Consider for example ak = −1, bk = 0. Comparison Test needs non-negative
terms.
P
(b) ak > 0 and kak → 0 implies ak converges.
FALSE: Take for example ak = 1/((k + 1) log(k + 1)).
P
(c) If ak is a convergent series of positive terms then 0 < ak < 1/k for k sufficiently
large.
FALSE: Take (
1/m2 if k = 2m for some m,
ak =
0 otherwise.
1/k 2 ). But, for
P P
Then ak converges (its partial sum sequence behaves like that for
k = 2m ,
2m
|kak | = 2 ,
m
and this is not bounded by 1 for large k—in fact it tends to infinity, as 8.9(b) shows.
Morals from (c):
(1) Series with gaps can be useful for counterexamples.
1/k p does not provide a counterexample here. It is naive to think that
P
(2) A series
behaviour of series of this special form is typical.
12.19. Postscript: other tests for convergence. Not every series you may meet can be
handled by applying directly one of the tests presented above. In particular you should be
aware that the Ratio Test is rather crude and quite often fails to give a result.
The list of tests for convergence, or for absolute convergence, that we have given is far
from exhaustive, and many other convergence tests exist. We note in particular
• Cauchy’s nth Root Test and Raabe’s Test—both useful alternatives to the Ratio
Test, the first because it works neatly on many power series (see Section 13) and the
second because it provides a backstop to the Ratio Test, giving a result in a number of
cases where the Ratio Test does not;
• Cauchy’s CondensationPTest Under the same conditions P mon f mas in the Integral
Test, the test asserts that f (k) converges if and only if 2 f (2 ) converges.
51
z 2k+1 z 2k
(−1)k and (−1)k
P P
• Sine and cosine: Consider . Applying the Ratio
(2k + 1)! (2k)!
Test (for z 6= 0) we can prove (do it for yourself!) that each of these series converges
absolutely for all z ∈ C and we define
∞ 2k+1 ∞
X
k z
X z 2k
sin z = (−1) and cos z = (−1)k .
k=0
(2k + 1)! k=0
(2k)!
52
—noting that we make the definitions legitimate by using the Ratio Test to prove that
the associated power series converge absolutely for all z ∈ C.
Note: We would not expect to be able to define tan, cot, sec and cosec, or their hyperbolic
analogues, as power series converging absolutely for all z ∈ C.
We’d like to work our way towards showing that, for a real variable anyway, these series
definitions do capture the properties we expect of the functions familiar from elementary
mathematics. We begin with some elementary facts.
13.3. AOL
P for series in general and power series in particular. Recall the key fact a
series
P a kPconverges if and only if its partial sum sequence (sn ) converges. Given two series
ak and bk with partial sum sequences (sn ) and (tn ), we have
(a1 + b1 ) + · · · + (an + bn ) = (a1 + · · · + an ) + (b1 + · · · + bn ) = sn + tn
P
(by properties of arithmetic—only
P a FINITE P number of terms involved!). Hence, if ak
convergesP
to s and bk converges
P to t, then (a k +b k ) converges, to s+t. Similar arguments
apply to (ak − bk ) and to cak , where c is a constant.
It follows in particular that we can add/subtract and scalar-multiply convergent power
series in the expected ‘term-by-term’ way.
1 1
cos z = (eiz + e−iz ), cosh z = (ez + e−z ),
2 2
1 1 z
sin z = (eiz − e−iz ), sinh z = (e − e−z ),
2i 2
One can also derive, straight from the definitions, the familiar Osborn’s Rules linking cos
and cosh and linking sin and sinh. For example, cos iz = cosh z, and this is valid for complex
z and not just when z is a real number.
(Remember that sup E (for E ⊆ R) exists if and only if E is non-empty and bounded above.
|ck 0k | converges, non-emptyness is not at issue here.)
P
Since
53
ck z k be
P
13.5. Proposition on radius of convergence (technical but important). Let
a power series with radius of convergence R (> 0). Then
|ck z k | converges for |z| < R, and hence ck z k converges for |z| < R.
P P
(i)
ck z k diverges if |z| > R.
P
(ii)
(ii) Assume for contradiction that there exists z with |z| > R such that ck z k converges.
P
Then ck z k → 0. Therefore (because a convergent sequence is bounded) there exists a constant
M such that |ck z k | 6 M . Pick ρ with |z| > ρ > R. Then
k ρ
k ρ k
k
0 6 |ck ρ | = |ck z | 6 M .
z z
k k
P P
But |ρ/z| is convergent, since |ρ/z| < 1. So |ck ρ | converges by comparison, and this
contradicts the definition of R.
Notes
• Part (i) of the proposition is not just the definition of R.
• Some sources replace |ck z k | by ck z k in the definition of R. The proposition implies
P P
that the two versions are equivalent. We prefer to phrase the definition with moduli in,
as a reminder that we can find R by using tests for convergence of series of non-negative
terms.
• We refer to { z ∈ C | |z| < R } as the disc of convergence. For real power series we
have an interval of convergence.
• If ck xk is a real power series with radius of convergence R then the series may converge
P
or diverge at R and −R, and similarly for |z| = R in the complex case.
(e) Problem sheet 6, Q. 4 is set as an exercise on convergence tests in the context of real
power series. With the definition of R in place it can be seen as asking you to calculate R
for various power series. Working through this exercise is intended to provide practice
in finding radius of convergence and to serve also to illustrate the properties of power
series captured by Proposition 13.5.
Important notes
• The last example shows that the Ratio Test cannot necessarily be used to find R because
the limit one needs to consider may not exist. More advanced texts give a formula for
R involving lim inf |ck+1 /ck |. Here lim inf is a notion which can sometimes be exploited
in place of lim when the latter fails to exist. This formula is not needed for Prelims,
and is frequently abused. DO NOT CLAIM that R = 1/ lim |ck+1 /ck | gives R for a
general power series.
If we are to make good use of power series as functions we would like to know they
have good behaviour, in particular that it is legitimate to differentiate them. The following
Differentiation Theorem is very important, but a first-principles proof is technical (you’ll see
this in Analysis II) and a better proof for real power series uses integration theory (Analysis
III). So we present the theorem here without proof.
ck xk be a real power
P
13.7. Differentiation Theorem for (real) power series. Let
series and assume that the series has radius of convergence R where 0 < R 6 ∞. Let
∞
X
f (x) := ck x k (|x| < R).
k=0
Then f (x) is well defined for each x with |x| < R and moreover the derivative f 0 (x) exists
for each such x and is given by
∞ ∞ ∞
d X ∗
Xd X
f 0 (x) = ck x k = ck x k = kck xk−1 .
dx k=0 k=0
dx k=1
Note: The theorem is very powerful and it is far from obvious that it is true. Note that
differentiation is carried out by a limiting process, that summing an infinite series also
involves a limiting process, and that iterated limits may not commute (recall Problem sheet
4, Points to Ponder B.). Saying we can differentiate term-by-term is exactly saying that we
can interchange the order of two limits: in ∗ above
∞ ∞
d X X d
ck x k ck x k
and
dx k=0 k=0
dx
13.8. Some applications of the (real) Differentiation Theorem. The general idea here
is to use the Differentiation Theorem to derive suitable differential equations whose solution
leads to formulae connecting functions defined by power series.
(a) Because the power series defining the functions are absolutely convergent for all real x,
from above, the Differentiation Theorem implies that the derivatives below exist and
are given by the expected formulae:
d x
e = ex ,
dx
d d
sin x = cos x, sinh x = cosh x,
dx dx
d d
cos x = − sin x, cosh x = sinh x.
dx dx
To illustrate:
∞ ∞ ∞ ∞
d x d X xk ∗ X d xk X kxk−1 X xm
e = = = = = ex .
dx dx k=0 k! k=0
dx k! k=1
k! m=0
m!
Here the equality marked ∗ holds by the Differentiation Theorem.
(d) Addition formulae for the trigonometric and hyperbolic functions are proved by adapt-
ing the strategy used for (c).
Note: You might be tempted to try to prove (b) by taking the power series for sin x and
squaring it, and likewise for cos x. Not recommended! Justifying multiplying power series
in the same way as one would multiply polynomials requires serious work. After all, you’ve
seen that infinite sums do not necessarily behave the same way as regards arithmetic as finite
sums do—beware · · · ! (See Problem sheet 7, Point to Ponder A. for more on this.)
56
13.9. Trigonometric functions: what became of π. You will probably have met the
cosine and sine series before, as Maclaurin expansions, but you will have been introduced
to the functions cos x and sin x geometrically. We have turned things around and used the
series to define these functions. One can capture the number known as π in either of the
following (equivalent) ways:
• π is the smallest number x > 0 for which sin x = 0;
• π/2 is the smallest number x > 0 for which cos x = 0.
[In Analysis II you’ll discover why these smallest zeros exist.] One can then use the addition
formulae to prove the periodicity results
cos(x + 2π) = cos x, sin(x + 2π) = sin x.
13.10. Looking ahead: the importance of complex power series. We have worked
with complex power series where the arguments are the same as for the real case. We have
restricted consideration of differentiation to the real case, because you have only learned
about differentiation of functions of a real variable so far. In the second year you will study
complex analysis, ‘complex’ meaning ‘in C’ and not ‘more complicated’.
One fact worth noting, and quite easy to prove from the definitions, is that
ex+iy = ex (cos y + i sin y) (for x, y ∈ R).
As a corollary: the facts about π then give
e2πi = 1.
[The remaining examples are on the boundary between Analysis I and Analysis II and III, and
won’t be covered in lectures. They provide additional illustrations of the use of differentiation
of power series to obtain valuable information about important functions.]
13.11. The binomial expansion. Here we shall freely use arbitrary powers of real num-
bers and formulae for their derivatives, noting that such powers are defined using exponentials
(and logs).
Let α ∈ R \ N. (The case that α ∈ N was covered in Introduction to University Mathe-
matics, and doesn’t involve convergence issues.) Consider the infinite sum
∞
X α(α − 1) . . . (α − k + 1)
Bα (x) = xk .
k=0
k!
We apply the Ratio Test to show the series has radius of convergence 1:
α(α − 1) . . . (α − k)xk+1 /(k + 1)!
α − k
lim
= lim
|x| = |x|.
α(α − 1) . . . (α − k + 1)xk /k! k +1
The Differentiation Theorem tells us we can differentiate term-by-term with respect to x for
|x| < 1, This gives
∞
X α(α − 1) . . . (α − k + 1)
Bα0 (x) = xk−1
k=1
(k − 1)!
∞
X α(α − 1) . . . (α − r) r
= x (writing k = r + 1).
r=0
r!
57
Now multiply by (1 + x), treating x as constant, and collect together coefficients associated
with the same power of x. We get
∞
0
X α(α − 1) . . . (α − s) α(α − 1) . . . (α − s + 1)
(1 + x)Bα (x) = + xs
s=0
s! (s − 1)!
∞
X α(α − 1) . . . (α − s + 1)
= ((α − s) + s)) xs .
s=0
s!
Therefore we have a differential equation for Bα :
(1 + x)Bα0 (x) = αBα (x).
Solving, we get Bα (x) = C(1 + x)α , where C is a constant (see Analysis II for justification).
Putting x = 0 we get C = 1.