Part IA - Analysis I: Based On Lectures by W. T. Gowers
Part IA - Analysis I: Based On Lectures by W. T. Gowers
Part IA - Analysis I: Based On Lectures by W. T. Gowers
Lent 2015
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Continuity
Continuity of real- and complex-valued functions defined on subsets of R and C. The
intermediate value theorem. A continuous function on a closed bounded interval is
bounded and attains its bounds. [3]
Differentiability
Differentiability of functions from R to R. Derivative of sums and products. The chain
rule. Derivative of the inverse function. Rolle’s theorem; the mean value theorem.
One-dimensional version of the inverse function theorem. Taylor’s theorem from R to
R; Lagrange’s form of the remainder. Complex differentiation. [5]
Power series
Complex power series and radius of convergence. Exponential, trigonometric and
hyperbolic functions, and relations between them. *Direct proof of the differentiability
of a power series within its circle of convergence*. [4]
Integration
Definition and basic properties of the Riemann integral. A non-integrable function.
Integrability of monotonic functions. Integrability of piecewise-continuous functions.
The fundamental theorem of calculus. Differentiation of indefinite integrals. Integration
by parts. The integral form of the remainder in Taylor’s theorem. Improper integrals. [6]
1
Contents IA Analysis I
Contents
0 Introduction 3
2 Convergence of sequences 7
2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Sums, products and quotients . . . . . . . . . . . . . . . . . . . . 8
2.3 Monotone-sequences property . . . . . . . . . . . . . . . . . . . . 10
2.4 Cauchy sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.5 Limit supremum and infimum . . . . . . . . . . . . . . . . . . . . 14
4 Continuous functions 23
4.1 Continuous functions . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Continuous induction* . . . . . . . . . . . . . . . . . . . . . . . . 26
5 Differentiability 29
5.1 Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2 Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.3 Differentiation theorems . . . . . . . . . . . . . . . . . . . . . . . 33
5.4 Complex differentiation . . . . . . . . . . . . . . . . . . . . . . . 37
2
0 Introduction IA Analysis I
0 Introduction
In IA Differential Equations, we studied calculus in a non-rigorous setting. While
we did define differentiation (properly) as a limit, we did not define what a limit
was. We didn’t even have a proper definition of integral, and just mumbled
something about it being an infinite sum.
In Analysis, one of our main goals is to put calculus on a rigorous foundation.
We will provide unambiguous definitions of what it means to take a limit, a
derivative and an integral. Based on these definitions, we will prove the results
we’ve had such as the product rule, quotient rule and chain rule. We will also
rigorously prove the fundamental theorem of calculus, which states that the
integral is the inverse operation to the derivative.
However, this is not all Analysis is about. We will study all sorts of limiting
(“infinite”) processes. We can see integration as an infinite sum, and differen-
tiation as dividing two infinitesimal quantities. In Analysis, we will also study
infinite series such as 1 + 14 + 19 + 16
1
+ · · · , as well as limits of infinite sequences.
Another important concept in Analysis is continuous functions. In some
sense, continuous functions are functions that preserve limit processes. While
their role in this course is just being “nice” functions to work with, they will be
given great importance when we study metric and topological spaces.
This course is continued in IB Analysis II, where we study uniform convergence
(a stronger condition for convergence), calculus of multiple variables and metric
spaces.
3
1 The real number system IA Analysis I
4
1 The real number system IA Analysis I
Definition (Least upper bound property). An ordered set X has the least upper
bound property if every non-empty subset of X that is bounded above has a
supremum.
Obvious modifications give rise to definitions of lower bound, greatest lower
bound (or infimum) etc. It is simple to check that an ordered field with the least
upper bound property has the greatest lower bound property.
Definition (Real numbers). The real numbers is an ordered field with the least
upper bound property.
Of course, it is very important to show that such a thing exists, or else we
will be studying nothing. It is also nice to show that such a field is unique (up
to isomorphism). However, we will not prove these in the course.
In a field, we can define the “natural numbers” to be 2 = 1 + 1, 3 + 1 + 2 etc.
Then an important property of the real numbers is
Lemma (Archimedean property v1)). Let F be an ordered field with the least
upper bound property. Then the set {1, 2, 3, · · · } is not bounded above.
Proof. If it is bounded above, then it has a supremum x. But then x − 1 is not
an upper bound. So we can find n ∈ {1, 2, 3, · · · } such that n > x − 1. But then
n + 1 > x, but x is supposed to be an upper bound.
Is the least upper bound property required to prove the Archimedean prop-
erty? It seems like any ordered field should satisfy this even if they do not have
the least upper bound property. However, it turns out there are ordered fields in
which the integers are bounded above.
P (x)
Consider the field of rational functions, i.e. functions in the form Q(x) with
P (x), Q(x) being polynomials, under the usual addition and multiplication. We
P (x) R(s)
order two functions Q(x) , S(x) as follows: these two functions intersect only
5
1 The real number system IA Analysis I
finitely many times because P (x)S(x) = R(x)Q(x) has only finitely many roots.
After the last intersection, the function whose value is greater counts as the
greater function. It can be checked that these form an ordered field.
In this field, the integers are the constant functions 1, 2, 3, · · · , but it is
bounded above since the function x is greater than all of them.
6
2 Convergence of sequences IA Analysis I
2 Convergence of sequences
Having defined real numbers, the first thing we will study is sequences. We
will want to study what it means for a sequence to converge. Intuitively, we
would like to say that 1, 12 , 13 , 14 , · · · converges to 0, while 1, 2, 3, 4, · · · diverges.
However, the actual formal definition of convergence is rather hard to get right,
and historically there have been failed attempts that produced spurious results.
2.1 Definitions
Definition (Sequence). A sequence is, formally, a function a : N → R (or C).
Usually (i.e. always), we write an instead of a(n). Instead of a, we usually write
it as (an ), (an )∞ ∞
1 or (an )n=1 to indicate it is a sequence.
(∃C)(∀n) |an | ≤ C.
7
2 Convergence of sequences IA Analysis I
8
2 Convergence of sequences IA Analysis I
6 0.
Lemma (Quotient of sequences). Let (an ) be a sequence such that (∀n) an =
Suppose that an → a and a 6= 0. Then 1/an → 1/a.
Proof. We have
1 1 a − an
− = .
an a aan
We want to show that this → 0. Since a − an → 0, we have to show that 1/(aan )
is bounded.
Since an → a, ∃N such that ∀n ≥ N , |an − a| ≤ a/2. Then ∀n ≥ N , |an | ≥
|a|/2. Then |1/(an a)| ≤ 2/|a|2 . So 1/(an a) is bounded. So (a − an )/(aan ) → 0
and the result follows.
Corollary. If an → a, bn → b, bn , b 6= 0. Then an /bn = a/b.
Proof. We know that 1/bn → 1/b. So the result follows by the product rule.
Lemma (Sandwich rule). Let (an ) and (bn ) be sequences that both converge to
a limit x. Suppose that an ≤ cn ≤ bn for every n. Then cn → x.
Proof. Let ε > 0. We can find N such that ∀n ≥ N , |an −x| < ε and |bn −x| < ε.
The ∀n ≥ N , we have x − ε < an ≤ cn ≤ bn < x + ε. So |cn − x| < ε.
Example. 1/2n → 0. For every n, n < 2n . So 0 < 1/2n < 1/n. The result
follows from the sandwich rule.
Example. We want to show that
n2 + 3 1
→ .
(n + 5)(2n − 1) 2
We can obtain this by
n2 + 3 1 + 3/n2 1
= → ,
(n + 5)(2n − 1) (1 + 5/n)(2 − 1/n) 2
by sum rule, sandwich rule, Archimedean property, product rule and quotient
rule.
Example. Let k ∈ N and let δ > 0. Then
nk
→ 0.
(1 + δ)n
This can be summarized as “exponential growth beats polynomial growth even-
tually”.
By the binomial theorem,
n n
(1 + δ) ≥ δ k+1 .
k+1
Also, for n ≥ 2k,
(n/2)k+1
n n(n − 1) · · · (n − k)
= ≥ .
k+1 (k + 1)! (k + 1)!
So for sufficiently large n,
nk nk 2k+1 (k + 1)! 2k+1 (k + 1)! 1
n
≤ k+1 k+1
= · → 0.
(1 + δ) n δ δ k+1 n
9
2 Convergence of sequences IA Analysis I
10
2 Convergence of sequences IA Analysis I
11
2 Convergence of sequences IA Analysis I
12
2 Convergence of sequences IA Analysis I
13
2 Convergence of sequences IA Analysis I
To see that this exists, set bn = supm≥n am . Then (bn ) is decreasing since we
are taking the supremum of fewer and fewer things, and is bounded below by
any lower bound for (an ) since bn ≥ an . So it converges.
Similarly, we define the limit infimum as
lim inf an = lim inf am .
n→∞ n→∞ m≥n
– an → a
– lim sup an = lim inf an = a.
Proof. If an → a, then let ε > 0. Then we can find an n such that
a − ε ≤ am ≤ a + ε for all m ≥ n
It follows that
a − ε ≤ inf am ≤ sup am ≤ a + ε.
m≥n m≥n
Conversely, if lim inf an = lim sup an = a, then let ε > 0. Then we can find n
such that
inf am > a − ε and sup am < a + ε.
m≥n m≥n
14
3 Convergence of infinite sums IA Analysis I
15
3 Convergence of infinite sums IA Analysis I
∞
X 1
Example. converges. This is since
n=2
n(n − 1)
1 1 1
= − .
n(n − 1) n−1 n
So
N
X 1 1
=1− → 1.
n=2
n(n − 1) N
16
3 Convergence of infinite sums IA Analysis I
∞
X n+5
(v) Consider . We show this converges by noting
n=1
n3 − 7n2 /2
7n2
7
n3 − = n2 n − .
2 2
So if n ≥ 8, then
7n2 n3
n3 − ≥ .
2 2
Also, n + 5 ≤ 2n. So
n+5
≤ 4/n2 .
n3 − 7n2 /2
So it converges by the comparison test.
1/nα converges.
P
(vi) If α > 1, then
PN
Let SN = n=1 1/nα . Then
1 1
S2n − S2n−1 = + ··· + n α
(2n−1 + 1)α (2 )
2n−1
≤ n−1 α
(2 )
= (2n−1 )1−α
= (21−α )n−1 .
It is easy to compare with 1/n2 to get that the partial sums S2n converges. But
S2n+1 − S2n = 1/(2n + 1) → 0, so the S2n+1 converges to the same limit.
It does not converge absolutely, because the sum of the absolute values is
the harmonic series.
17
3 Convergence of infinite sums IA Analysis I
P P
Lemma. Let an converge absolutely. Then an converges.
P PN PN
Proof. We know that |an | converges. Let Sn = n=1 an and TN = n=1 |an |.
We know two ways to show random sequences converge, without knowing
what they converge to, namely monotone-sequences and Cauchy sequences. Since
Sn is not monotone, we shall try Cauchy sequences.
If p > q, then
p p
X X
|Sp − Sq | = an ≤ |an | = Tp − Tq .
q=1 n=q+1
But the sequence Tp converges. So ∀ε > 0, we can find N such that for all
p > q ≥ N , we have Tp − Tq < ε, which implies |Sp − Sq | < ε.
P
Definition (Unconditional
P∞ convergence). A series an converges uncondition-
ally if the series n=1 aπ(n) converges for every bijection π : N → N, i.e. no
matter how we re-order the elements of an , the sum still converges.
P
Theorem. If an converges absolutely, then it converges unconditionally.
PN
Proof. Let Sn = n=1 aπ(n) . Then if p > q,
p ∞
X X
|Sp − Sq | = aπ(n) ≤ |aπ(n) |.
n=q+1 n=q+1
P P∞
Let ε > 0. Since |an | converges, pick M such that n=M +1 |an | < ε.
Pick N large enough that {1, · · · , M } ⊆ {π(1), · · · , π(N )}.
Then if n > N , we have π(n) > M . Therefore if p > q ≥ N , then
p
X ∞
X
|Sp − Sq | ≤ |aπ(n) | ≤ |an | < ε.
n=q+1 n=M +1
18
3 Convergence of infinite sums IA Analysis I
and
{π(1), · · · , π(M )} ⊆ {1, · · · , N }.
Then for every K ≥ N ,
K K
K K
X X X X ε ε
an − aπ(n) ≤ |an | + |aπ(n) | < + = ε.
2 2
n=1 n=1 n=M +1 n=M +1
19
3 Convergence of infinite sums IA Analysis I
Example.
1 1 1
1− √
3
+ √
3
− √
3
+ · · · converges.
2 3 4
Lemma (Ratio test). We have three versions:
|an+1 |
(∀n ≥ N ) ≤ c,
|an |
P
then an converges. Note that just because the ratio is always less than
1, it doesn’t necessarily converge. It has to
Pbe always less than a fixed
number c. Otherwise the test will say that 1/n converges.
Proof.
≤ cn−1 |a1 |. Since
P n P
(i) |an |P c converges, so does |an | by comparison test.
So an converges absolutely, so it converges.
|aN +k | ≤ ck |aN |. So the series
P
(ii) For all k ≥ 0, we have P |aN +k | converges,
and therefore so does |ak |.
an+1 |an+1 |
(iii) If an → ρ, then |an | → |ρ|. So (setting ε = (1 − |ρ|)/2) there exists N
such that ∀n ≥ N , |a|an+1
n|
|
≤ 1+|ρ|
2 < 1. So the result follows from (ii).
(n + 1)bn+1
an+1 1
= = 1 + b → b < 1.
an nbn n
So it converges.
∞ X
X ∞
We can also evaluate this directly by considering bn .
i=1 n=i
The following two tests were taught at the end of the course, but are included
here for the sake of completeness.
20
3 Convergence of infinite sums IA Analysis I
Theorem
P∞ (Condensation test). Let (an ) be a decreasing non-negative sequence.
Then n=1 an < ∞ if and only if
∞
X
2k a2k < ∞.
k=1
P1 P 1
Proof. This is basically the proof that n diverges and nα converges for
α < 1 but written in a more general way.
We have
a1 + (a2 + a3 ) + (a4 + · · · + a7 ) + · · ·
≤a1 + 2a2 + 4a4 + · · ·
1
P∞ P∞ 1
Example. If an = n, then 2k a2k = 1. So k=1 2k a2k = ∞. So n=1 n = ∞.
After we formally define integrals, we will prove the integral test:
21
3 Convergence of infinite sums IA Analysis I
So it is Cauchy. So it converges
an (z n+1 − z n ) into aN z N +1 −
P
Note Pthat here we transformed the sum
M n+1
aM z + (an −an+1 )z . What we have effectively done is a discrete analogue
of integrating by parts.
P n
Example. The series z /n converges if |z| < 1 or if |z| = 1 and z 6= 1, and
it diverges if z = 1 or |z| > 1.
The cases |z| < 1 and |z| > 1 are trivial from the ratio test, and Abel’s test
is required for the |z| = 1 cases.
22
4 Continuous functions IA Analysis I
4 Continuous functions
4.1 Continuous functions
Definition (Continuous function). Let A ⊆ R, a ∈ A, and f : A → R. Then f
is continuous at a if for any ε > 0, there is some δ > 0 such that if y ∈ A is such
that |y − a| < δ, then |f (y) − f (a)| < ε. In symbols, we have
(∀a ∈ A)(∀ε > 0)(∃δ > 0)(∀y ∈ A) |y − a| < δ ⇒ |f (y) − f (a)| < ε.
Suppose that we don’t know what the function actually is, but we have a
computer program that computes this function. We want to know what f (π)
is. Since we cannot input π (it has infinitely many digits), we can try 3, and it
gives 0. Then we try 3.14, and it gives 0 again. If we try 3.1416, it gives 1 (since
π = 3.1415926 · · · < 3.1416). We keep giving more and more digits of π, but the
result keeps oscillating between 0 and 1. We have no hope of what f (π) might
be, even approximately. So this f is discontinuous at π.
However, if we have the function g(x) = x2 , then we can find the (approx-
imate) value of g(π). We can first try g(3) and obtain 9. Then we can try
g(3.14) = 9.8596, g(3.1416) = 9.86965056 etc. We can keep trying and obtain
more and more accurate values of g(π). So g is continuous at π.
Example.
– Constant functions are continuous.
– f is continuous
– If (an ) is a sequence in A with an → a, then f (an ) → f (a).
Proof. (i)⇒(ii) Let ε > 0 Since f is continuous at a,
23
4 Continuous functions IA Analysis I
For each n, we can therefore pick an ∈ A such that |an − a| < n1 and |f (an ) −
f (a)| ≥ ε. But then an → a (by Archimedean property), but f (an ) 6→ f (a).
Example.
(
−1 x < 0
(i) Let f (x) = . Then f is not continuous because − n1 → 0 but
1 x≥0
f (− n1 ) → −1 6= f (0).
(ii) Let f : Q → R with (
1 x2 > 2
f (x) =
0 x2 < 2
Then f is continuous. For every a ∈ Q, we can find an interval about a on
which f is constant. So f is continuous at a.
(iii) Let (
sin x1 x 6= 0
f (x) =
0 x=0
Then f (a) is discontinuous. For example, let an = 1/[(2n + 0.5)π]. Then
an → 0 and f (an ) → 1 6= f (0).
We can use this sequence definition as the definition for continuous functions.
This has the advantage of being cleaner to write and easier to work with. In
particular, we can reuse a lot of our sequence theorems to prove the analogous
results for continuous functions.
Lemma. Let A ⊆ R and f, g : A → R be continuous functions. Then
(i) f + g is continuous
(ii) f g is continuous
(iii) if g never vanishes, then f /g is continuous.
Proof.
(i) Let a ∈ A and let (an ) be a sequence in A with an → a. Then
24
4 Continuous functions IA Analysis I
With this lemma, from the fact that constant functions and f (x) = x are
continuous, we know that all polynomials are continuous. Similarly, rational
functions P (x)/Q(x) are continuous except when Q(x) = 0.
Lemma. Let A, B ⊆ R and f : A → B, g : B → R. Then if f and g are
continuous, g ◦ f : A → R is continuous.
Proof. We offer two proofs:
25
4 Continuous functions IA Analysis I
It is easy to generalize this to get that, if f (a) < c < f (b), then ∃x ∈ (a, b)
such that f (x) = c, by applying the result to f (x) − c. Also, we can assume
instead that f (b) < c < f (a) and obtain the same result by looking at −f (x).
Corollary. Let f : [a, b] → [c, d] be a continuous strictly increasing function
with f (a) = c, f (b) = d. Then f is invertible and its inverse is continuous.
Proof. Since f is strictly increasing, it is an injection (suppose x 6= y. wlog,
x < y. Then f (x) < f (y) and so f (x) 6= f (y)). Now let y ∈ (c, d). By the
intermediate value theorem, there exists x ∈ (a, b) such that f (x) = y. So f is a
surjection. So it is a bijection and hence invertible.
Let g be the inverse. Let y ∈ [c, d] and let ε > 0. Let x = g(y). So f (x) = y.
Let u = f (x − ε) and v = f (x + ε) (if y = c or d, make the obvious adjustments).
Then u < y < v. So we can find δ > 0 such that (y − δ, y + δ) ⊆ (u, v). Then
|z − y| < δ ⇒ g(z) ∈ (x − ε, x + ε) ⇒ |g(z) − g(y)| < ε.
√
With this corollary, we can create more continuous functions, e.g. x.
Proposition (Continuous induction v1). Let a < b and let A ⊆ [a, b] have the
following properties:
(i) a ∈ A
(ii) If x ∈ A and x 6= b, then ∃y ∈ A with y > x.
(ii) If [a, x] ⊆ A and x 6= b, then there exists y > x such that [a, y] ⊆ A.
26
4 Continuous functions IA Analysis I
If x 6∈ A, then f (x) > 0 (we assume that f is never zero. If not, we’re done).
Then by continuity, ∃δ > 0 such that |y − x| < δ ⇒ f (y) > 0. So y ∈
6 A.
Hence by continuous induction, b ∈ A. Contradiction.
Now we prove that continuous functions in closed intervals are bounded.
Theorem. Let [a, b] be a closed interval in R and let f : [a, b] → R be continuous.
Then f is bounded.
Proof. Let f : [a, b] be continuous. Let A = {x : f is bounded on [a, x]}. Then
a ∈ A. If x ∈ A, x 6= b, then ∃δ > 0 such that |y − x| < δ ⇒ |f (y) − f (x)| < 1.
So ∃y > x (e.g. take min{x + δ/2, b}) such that f is bounded on [a, y], which
implies that y ∈ A.
Now suppose that ∀ε > 0, ∃y ∈ (x, −ε, x] such that y ∈ A. Again, we can
find δ > 0 such that f is bounded on (x − δ, x + δ), and in particular on (x − δ, x].
Pick y such that f is bounded on [a, y] and y > x − δ. Then f is bounded on
[a, x]. So x ∈ A.
So we are done by continuous induction.
Finally, we can prove a theorem that we have not yet proven.
Definition (Cover of a set). Let A ⊆ R. A cover of A bySopen intervals is a
set {Iγ : γ ∈ Γ} where each Iγ is an open interval and A ⊆ γ∈Γ Iγ .
A finite subcover is a finite subset {Iγ1 , · · · , Iγn } of the cover that is still a
cover.
Not every cover has a finite subcover. For example, the cover {( n1 , 1) : n ∈ N}
of (0, 1) has no finite subcover.
27
4 Continuous functions IA Analysis I
28
5 Differentiability IA Analysis I
5 Differentiability
In the remainder of the course, we will properly develop calculus, and put
differentiation and integration on a rigorous foundation. Every notion will be
given a proper definition which we will use to prove results like the product and
quotient rule.
5.1 Limits
First of all, we need the notion of limits. Recall that we’ve previously had limits
for sequences. Now, we will define limits for functions.
Definition (Limit of functions). Let A ⊆ R and let f : A → R. We say
lim f (x) = `,
x→a
or “f (x) → ` as x → a”, if
We couldn’t care less what happens when x = a, hence the strict inequality
0 < |x − a|. In fact, f doesn’t even have to be defined at x = a.
Example. Let (
x x 6= 2
f (x) =
3 x=2
Then lim = 2, even though f (2) = 3.
x→2
sin x
Example. Let f (x) = x . Then f (0) is not defined but lim f (x) = 1.
x→0
We will see a proof later after we define what sin means.
5.2 Differentiation
Similar to what we did in IA Differential Equations, we define the derivative as
a limit.
29
5 Differentiability IA Analysis I
Proposition.
f (x + h) = f (x) + hf 0 (x) + o(h).
We can interpret this as an approximation of f (x + h):
f (x + h) = f (x) + hf 0 (x) + o(h) .
| {z } |{z}
linear approximation error term
And differentiability shows that this is a very good approximation with small
o(h) error.
Conversely, we have
Proposition. If f (x + h) = f (x) + hf 0 (x) + o(h), then f is differentiable at x
with derivative f 0 (x).
Proof.
f (x + h) − f (x) o(h)
= f 0 (x) + → f 0 (x).
h h
30
5 Differentiability IA Analysis I
We can prove the usual rules of differentiation using the small o-notation. It
can also be proven by considering limits directly, but the notation will become a
bit more daunting.
Lemma (Sum and product rule). Let f, g be differentiable at x. Then f + g
and f g are differentiable at x, with
Proof.
(f + g)(x + h) = f (x + h) + g(x + h)
= f (x) + hf 0 (x) + o(h) + g(x) + hg 0 (x) + o(h)
= (f + g)(x) + h(f 0 (x) + g 0 (x)) + o(h)
f g(x + h) = f (x + h)g(x + h)
= [f (x) + hf 0 (x) + o(h)][g(x) + hg 0 (x) + o(h)]
= f (x)g(x) + h[f 0 (x)g(x) + f (x)g 0 (x)]
+ o(h)[g(x) + f (x) + hf 0 (x) + hg 0 (x) + o(h)] + h2 f 0 (x)g 0 (x)
| {z }
error term
We want to show that the error term is o(h), i.e. it divided by h tends to 0 as
h → 0.
But ε1 (h)g 0 (f (x)) → 0, f 0 (x)+ε1 (h) is bounded, and ε2 (hf 0 (x)+hε1 (h)) → 0
because hf 0 (x) + hε1 (h) → 0 and ε2 (0) = 0. So our error term is o(h).
We usually don’t write out the error terms so explicitly, and just use heuristics
like f (x + o(h)) = f (x) + o(h); o(h) + o(h) = o(h); and g(x) · o(h) = o(h) for any
(bounded) function g.
31
5 Differentiability IA Analysis I
Example.
by limit theorems.
Lemma (Quotient rule). If f and g are differentiable at x, and g(x) 6= 0, then
f /g is differentiable at x with derivative
0
f f 0 (x)g(x) − g 0 (x)f (x)
(x) = .
g g(x)2
Proof. First note that 1/g(x) = h(g(x)) where h(y) = 1/y. So 1/g(x) is differen-
−1 0
tiable at x with derivative g (x) by the chain rule.
g(x)2
By the product rule, f /g is differentiable at x with derivative
32
5 Differentiability IA Analysis I
g(y + k) − g(y)
→ [f 0 (x)]−1 = [f 0 (g(y)]−1 .
k
Example. Let f (x) = x1/2 for x > 0. Then f is the inverse of g(x) = x2 . So
1 1 1
f 0 (x) = = 1/2 = x−1/2 .
g 0 (f (x)) 2x 2
It is intuitively obvious: if you move up and down, and finally return to the
same point, then you must have changed direction some time. Then f 0 (x) = 0
at that time.
Proof. If f is constant, then we’re done.
Otherwise, there exists u such that f (u) 6= f (a). wlog, f (u) > f (a). Since f
is continuous, it has a maximum, and since f (u) > f (a) = f (b), the maximum
is not attained at a or b.
Suppose maximum is attained at x ∈ (a, b). Then for any h 6= 0, we have
(
f (x + h) − f (x) ≤ 0 h > 0
h ≥0 h<0
33
5 Differentiability IA Analysis I
Corollary (Mean value theorem). Let f be continuous on [a, b] (a < b), and
differentiable on (a, b). Then there exists x ∈ (a, b) such that
f (b) − f (a)
f 0 (x) = .
b−a
f (b)−f (a)
Note that b−a is the slope of the line joining f (a) and f (b).
f (x) f (b)
f (a)
The mean value theorem is sometimes described as “rotate your head and
apply Rolle’s”. However, if we actually rotate it, we might end up with a
non-function. What we actually want is a shear.
Proof. Let
f (b) − f (a)
g(x) = f (x) − x.
b−a
Then
f (b) − f (a)
g(b) − g(a) = f (b) − f (a) − (b − a) = 0.
b−a
So by Rolle’s theorem, we can find x ∈ (a, b) such that g 0 (x) = 0. So
f (b) − f (a)
f 0 (x) = ,
b−a
as required.
We’ve always assumed that if a function has a positive derivative everywhere,
then the function is increasing. However, it turns out that this is really hard to
prove directly. It does, however, follow quite immediately from the mean value
theorem.
Example. Suppose f 0 (x) > 0 for every x ∈ (a, b). Then for u, v in [a, b], we can
find w ∈ (u, v) such that
f (v) − f (u)
= f 0 (w) > 0.
v−u
It follows that f (v) > f (u). So f is strictly increasing.
Similarly, if f 0 (x) ≥ 2 for every x and f (0) = 0, then f (1) ≥ 2, or else we
can find x ∈ (0, 1) such that
f (1) − f (0)
2 ≤ f 0 (x) = = f (1).
1−0
34
5 Differentiability IA Analysis I
theorem.
So the result follows by induction.
Corollary. Suppose that f and g are both differentiable on an open interval
containing [a, b] and that f (k) (a) = g (k) (a) for k = 0, 1, · · · , n − 1, and also
f (b) = g(b). Then there exists x ∈ (a, b) such that f (n) (x) = g (n) (x).
Proof. Apply generalised Rolle’s to f − g.
Now we shall show that for any f , we can find a polynomial p of degree at
most n that satisfies the conditions for g, i.e. a p such that p(k) (a) = f (k) (a) for
k = 0, 1, · · · , n − 1 and p(b) = f (b).
A useful ingredient is the observation that if
(x − a)k
Qk (x) = ,
k!
then (
(j) 1 j=k
Qk (a) =
0 j 6= k
Therefore, if
n−1
X
Q(x) = f (k) (a)Qk (x),
k=0
35
5 Differentiability IA Analysis I
then
Q(j) (a) = f (j) (a)
for j = 0, 1, · · · , n − 1. To get p(b) = f (b), we use our nth degree polynomial
term:
(x − a)n
p(x) = Q(x) + n
f (b) − Q(b) .
(b − a)
Then our final term does not mess up our first n − 1 derivatives, and gives
p(b) = f (b).
By the previous corollary, we can find x ∈ (a, b) such that
That is,
n!
f (n) (x) =
f (b) − Q(b) .
(b − a)n
Therefore
(b − a)n (n)
f (b) = Q(b) + f (x).
n!
Alternatively,
(b − a)n−1 (n−1) (b − a)n (n)
f (b) = f (a) + (b − a)f 0 (a) + · · · + f (a) + f (x).
(n − 1)! n!
Setting b = a + h, we can rewrite this as
Theorem (Taylor’s theorem with the Lagrange form of remainder).
While it seems like we can prove this works by differentiating it and see that
f 0 (x) = f (x), the sum rule only applies for finite sums. We don’t know we can
differentiate a sum term by term. So we have to use Taylor’s theorem.
36
5 Differentiability IA Analysis I
Since f 0 (x) = f (x), it follows that all derivatives exist. By Taylor’s theorem,
If this seems weird, this is because we often think of C as R2 , but they are
not the same. For example, reflection is a linear map in R2 , but not in C. A
linear map in C is something in the form x 7→ bx, which can only be a dilation
or rotation, not reflections or other weird things.
Example. f (z) = |z| is also not differentiable. If it were, then |z|2 would be as
2
well (by the product rule). So would |z|z = z̄ when z = 6 0 by the quotient rule.
At z = 0, it is certainly not differentiable, since it is not even differentiable on R.
37
6 Complex power series IA Analysis I
|an z n | ≤ C
38
6 Complex power series IA Analysis I
∞
X zn
Example. has radius of convergence 1, since the ratio of (n + 1)th term
n=0
n
to nth is
z n+1 /(n + 1) n
n
=z· → z.
z /n n+1
So if |z| < 1, then the series converges by the ratio test. If |z| > 1, then eventually
the terms are increasing in modulus.
If z = 1, then it diverges (harmonic series). If |z| = 1 and z = 6 1, it converges
by Abel’s test.
∞
X zn
Example. The series converges for |z| ≤ 1 and diverges for |z| > 1.
n=1
n2
As evidenced by the above examples, the ratio test can be used to find the
radius of convergence. We also have an alternative test based on the nth root.
an z n is
P
Lemma. The radius of convergence of a power series
1
R= p .
lim sup n
|an |
p
Often n
|an | converges, so we only have to find the limit.
p p
Proof. Suppose |z| < 1/ lim sup n |an |. Then |z| lim sup n |an | < 1. Therefore
there exists N and ε > 0 such that
p
sup |z| n |an | ≤ 1 − ε
n≥N
|an z n | ≤ (1 − ε)n
for
P every n ≥ N , which implies (by comparison with geometric series) that
an z n converges absolutely. p p
On the other hand, if |z| lim sup n |an | > 1, it follows that |z| nP|an | ≥ 1 for
infinitely many n. Therefore |an z n | ≥ 1 for infinitely many n. So an z n does
not converge.
zn p 1
Example. The radius of convergence of n is 2 because n |an | = 2 for every
p p2
n. So lim sup n |an | = 12 . So 1/ lim sup n |an | = 2.
But often it is easier to findP
the radius convergence from elementary methods
such as the ratio test, e.g. for n2 z n .
39
6 Complex power series IA Analysis I
ez+w = ez ew .
ez+h − ez eh − 1
= ez
h h
h2
h
= ez 1 + + + ···
2! 3!
But
2 2 3
+ h + · · · ≤ |h| + |h| + |h| + · · · = |h|/2 → 0.
h
2! 3! 2 4 8 1 − |h|/2
So
ez+h − ez
→ ez .
h
n
and equate coefficients of z , you get
cn = a0 bn + a1 bn−1 + a2 bn−2 + · · · + an b0 .
P∞ P∞
Theorem. Let n=0 an and n=0 bn be two absolutely convergent series, P∞ and
let (cn ) be the convolution of the sequences (an ) and (bn ). Then n=0 cn
converges (absolutely), and
∞ ∞
! ∞ !
X X X
cn = an bn .
n=0 n=0 n=9
P
Proof. We first show that a rearrangement of cn converges absolutely.
P Hence
it converges unconditionally, and we can rearrange it back to cn .
Consider the series
Let
N
X N
X N
X N
X
SN = an , TN = bn , UN = |an |, VN |bn |.
n=0 n=0 n=0 n=0
40
6 Complex power series IA Analysis I
P P
Also let SN → S, TN → T, UN → U, VN → V (these exist since an and bn
converge absolutely).
If we take the modulus of the terms of (∗), and consider the first (N + 1)2
terms (i.e. the first N + 1 brackets), the sum is UN VN . Hence the series converges
absolutely to U V . Hence (∗) converges.
The partial sum up to (N + 1)2 of the series (∗) itself is SN TN , which
converges to ST . So the whole series converges to ST .
Since it converges absolutely, it converges unconditionally. Now consider a
rearrangement:
a0 b0 + (a0 b1 + a1 b0 ) + (a0 b2 + a1 b1 + a2 b0 ) + · · ·
Then this converges to ST as well. But the partial sum of the first 1 + 2 + · · · + N
terms is c0 + c1 + · · · + cN . So
∞
N
! ∞ !
X X X
cn → ST = an bn .
n=0 n=0 n=9
Corollary.
ez ew = ez+w .
Proof. By theorem above (and definition of ez ),
∞
wn z wn−1 z 2 wn−2 zn
X
z w
e e = 1· + + + ··· + ·1
n=0
n! 1! (n − 1)! 2! (n − 2)! n!
∞
z w
X 1 n n n−1 n 2 n−2 n n
e e = w + zw + z w + ··· + z
n=0
n! 1 2 n
∞
X
= (z + w)n by the binomial theorem
n=0
= ez+w .
eiz − e−iz z3 z5 z7
sin z = =z− + − + ···
2i 3! 5! 7!
eiz + e−iz z2 z4 z6
cos z = =1− + − + ··· .
2 2! 4! 6!
We now prove certain basic properties of sin and cos, using known properties
of ez .
41
6 Complex power series IA Analysis I
Proposition.
d ieiz + ie−iz
sin z = = cos z
dz 2i
d ieiz − ie−iz
cos z = = − sin z
dz 2
−2iz
2iz
e +2+e e2iz − 2 + e−2iz
sin2 z + cos2 z = + = 1.
4 −4
It follows that if x is real, then | cos x| and | sin x| are at most 1.
Proposition.
Proof.
So
sin(z + w) = sin z cos w + cos z sin w.
d
When x is real, we know that cos x ≤ 1. Also sin 0 = 0, and dx sin x =
cos x ≤ 1. So for x ≥ 0, sin x ≤ x, “by the mean value theorem”. Also, cos 0 = 1,
d
and dx cos x = − sin x, which, for x ≥ 0, is greater than −x. From this, it follows
2 2
that when x ≥ 0, cos x ≥ 1 − x2 (the 1 − x2 comes from “integrating” −x, (or
finding a thing whose derivative is −x)).
Continuing in this way, we get that for x ≥ 0, if you take truncate the power
series for sin x or cos x, it will be ≥ sin x, cos x if you stop at a positive term,
and ≤ if you stop at a negative term. For example,
x3 x5 x7 x9 x11
sin x ≥ x − + − + − .
3! 5! 7! 9! 11!
In particular,
22 24 2
cos 2 ≤ 1 − + = 1 − 2 + < 0.
2! 4! 3
Since cos 0 = 1, it follows by the intermediate value theorem that there exists
2
some x ∈ (0, 2) such that cos x = 0. Since cos x ≥ 1 − x2 , we can further deduce
that x > 1.
π
Definition (Pi). Define the smallest x such that cos x = 0 to be 2.
42
6 Complex power series IA Analysis I
Since sin2 z + cos2 z = 1, it follows that sin π2 = ±1. Since cos x > 0 on [0, π2 ],
sin π2 ≥ 0 by the mean value theorem. So sin π2 = 1.
Thus
Proposition.
π
cos z + = − sin z
2
π
sin z + = cos z
2
cos(z + π) = − cos z
sin(z + π) = − sin z
cos(z + 2π) = cos z
sin(z + 2π) = sin z
Proof.
π π π
cos z + = cos z cos − sin z sin
2 2 2
π
= − sin z sin
2
= − sin z
and similarly for others.
43
6 Complex power series IA Analysis I
converges absolutely.
Proof. Apply Lemma above again and divide by 2.
an z n be a power series with radius of convergence R. For
P
Theorem. Let
|z| < R, let
X∞ ∞
X
f (z) = an z n and g(z) = nan z n−1 .
n=0 n=1
Then f is differentiable with derivative g.
Proof. We want f (z + h) − f (z) − hg(z) to be o(h). We have
∞
X
f (z + h) − f (z) − hg(z) = an ((z + h)n − z n − hnz n ).
n=2
We want the huge infinite series to be bounded, and then the whole thing is a
bounded thing times h2 , which is definitely o(h).
Pick r such that |z| < r < R. If h is small enough that |z + h| ≤ r, then the
last infinite series is bounded above (in modulus) by
∞ ∞
X X n n−2
|an |(rn−2 + 2rn−2 + · · · + (n − 1)rn−2 ) = |an | r ,
n=2 n=2
2
44
6 Complex power series IA Analysis I
In IB Analysis II, we will prove the same result using the idea of uniform
convergence, which gives a much nicer proof.
Example. The derivative of
z2 z3
ez = 1 + z + + + ···
2! 3!
is
z2
1+z+ + · · · = ez .
2!
So we have another proof that of this fact.
Similarly, the derivatives of sin z and cos z work out as cos z and − sin z.
ez + e−z z2 z4 z6
cosh z = =1+ + + + ···
2 2! 4! 6!
ez − e−z z3 z5 z7
sinh z = =z+ + + + ···
2 3! 5! 7!
Either from the definition or from differentiating the power series, we get
that
Proposition.
d
cosh z = sinh z
dz
d
sinh z = cosh z
dz
Also, by definition, we have
Proposition.
cosh iz = cos z
sinh iz = i sin z
Also,
Proposition.
cosh2 z − sinh2 z = 1,
45
7 The Riemann Integral IA Analysis I
Xn
LD (f ) = (xi − xi−1 ) inf f (x)
x∈[xi−1 ,xi ]
i=1
The upper sum is the total area of the red rectangles, while the lower sum is
the total area of the black rectangles:
y
···
···
x
a x1 x2 x3 xi xi+1 · · · b
46
7 The Riemann Integral IA Analysis I
Using the picture above, this is because if we cut up the dissections into
smaller pieces, the red rectangles can only get chopped into shorter pieces and
the black rectangles can only get chopped into taller pieces.
y y
x x
x0 x1 x0 x1 x2 x3
But supx∈[xi−1 ,z] f (x) and supx∈[z,xi ] f (x) are both at most Mi . So this is at
most Mi (z − xi−1 + xi − z − (xi − xi−1 )) = 0. So
UD2 f ≤ UD1 f.
UD1 f ≥ LD2 f.
Proof. Let D be the least common refinement (or indeed any common refinement).
Then by lemma above (and by definition),
UD1 f ≥ UD f ≥ LD ≥ LD2 f.
47
7 The Riemann Integral IA Analysis I
If these are equal, then we call their common value the Riemann integral of f ,
Rb
and is denoted a f (x) dx.
If this exists, we say f is Riemann integrable.
We will later prove the fundamental theorem of calculus, which says that
integration is the reverse of differentiation. But why don’t we simply define
integration as anti-differentiation, and prove that it is the area of the curve?
There are things that we cannot find (a closed form of) the anti-derivative of,
2
like e−x . In these cases, we wouldn’t want to say the integral doesn’t exist — it
surely does according to this definition!
There is an immediate necessary condition for Riemann integrability — bound-
edness. If f is unbounded above in [a, b], then for any dissection D, there must
be some i such that f is unbounded on [xi−1 , xi ]. So Mi = ∞. So UD f = ∞.
Similarly, if f is unbounded below, then LD f = −∞. So unbounded functions
are not Riemann integrable.
Example. Let f (x) = x on [a, b]. Intuitively, we know that the integral is
(b2 − a2 )/2, and we will show this using the definition above. Let D = x0 < x1 <
· · · < xn be a dissection. Then
n
X
UD f = (xi − xi−1 )xi
i=1
b2 −a2
We know that the integral is 2 . So we put each term of the sum into the
x2i −x2i−1
form 2 plus some error terms.
n 2
x2i−1 x2 x2
X x
= i
− + i − xi−1 xi + i−1
i=1
2 2 2 2
n
1X 2
= (x − x2i−1 + (xi − xi−1 )2 )
2 i=1 i
n
1 2 1X
= (b − a2 ) + (xi − xi−1 )2 .
2 2 i=1
48
7 The Riemann Integral IA Analysis I
Similarly,
Z b
1 2
x dx > (b − a2 ) − ε.
a 2
So Z b
1 2
x dx = (b − a2 ).
a 2
Example. Define f : [0, 1] → R by
(
1 x∈Q
f (x) = .
0 x 6∈ Q
UD − LD < ε.
49
7 The Riemann Integral IA Analysis I
UD f − LD f < ε,
then
inf UD f − sup LD f < ε,
which is, by definition, that
Z b Z b
f (x) dx − f (x) dx < ε.
a a
So f is integrable.
The next big result we want to prove is that integration is linear, ie
Z b Z b Z b
(λf (x) + µg(x)) dx = λ f (x) dx + µ g(x) dx.
a a a
UD (λf ) = λUD f
LD (λf ) = λLD f.
50
7 The Riemann Integral IA Analysis I
Therefore
n
X
UD (−f ) = (xi − xi−1 )(−mi ) = −LD (f ).
i=1
Similarly,
LD (−f ) = −UD f.
So
UD (−f ) − LD (−f ) = UD f − LD f.
Hence if f is integrable, then −f is integrable by the Riemann integrability
criterion.
Proposition. Let f, g : [a, b] → R be integrable. Then f + g is integrable, and
Z b Z b Z b
(f (x) + g(x)) dx = f (x) dx + g(x) dx.
a a a
= UD f + UD g
Therefore,
Z b Z b Z b Z b Z b
(f (x) + g(x)) dx ≤ f (x) dx + g(x) dx = f (x) dx + g(x) dx.
a a a a a
Similarly,
Z b Z b Z b
(f (x) + g(x)) dx ≥ f (x) dx + g(x) dx.
a a a
So the upper and lower integrals are equal, and the result follows.
So we now have that
Z b Z b Z b
(λf (x) + µg(x)) dx = λ f (x) dx + µ g(x) dx.
a a a
51
7 The Riemann Integral IA Analysis I
Similarly,
For any pair of real numbers, x, y, we have that ||x| − |y|| ≤ |x − y| by the
triangle inequality. Then for any interval u, v ∈ [xi−1 , xi ], we have
Hence we have
UD (|f |) − LD (|f |) ≤ UD (f ) − LD (f ).
|f (x) − g(x)| ≤ C,
Similarly, if f is integrable on [a, b] and [b, c], then it is integrable on [a, c] and
the above equation also holds.
Proof. Let ε > 0, and let a = x0 < x1 < · · · < xn = c be a dissection of D of
[a, c] such that Z c
UD (f ) ≤ f (x) dx + ε,
a
and Z c
LD (f ) ≥ f (x) dx − ε.
a
52
7 The Riemann Integral IA Analysis I
Let D0 be the dissection made of D plus the point b. Let D1 be the dissection of
[a, b] made of points of D0 from a to b, and D2 be the dissection of [b, c] made of
points of D0 from b to c. Then
UD1 (f ) + UD2 (f ) = UD0 (f ) ≤ UD (f ),
and
LD1 (f ) + LD2 (f ) = LD0 (f ) ≥ LD (f ).
Since UD (f ) − LD (f ) < 2ε, and both UD2 (f ) − LD2 (f ) and UD1 (f ) − LD1 (f )
are non-negative, we have UD1 (f ) − LD1 (f ) and UD2 (f ) − LD2 (f ) are less than
2ε. Since ε is arbitrary, it follows that the restrictions of f to [a, b] and [b, c] are
both Riemann integrable. Furthermore,
Z b Z c Z c
f (x) dx+ f (x) dx ≤ UD1 (f )+UD2 (f ) = UD0 (f ) ≤ UD (f ) ≤ f (x) dx+ε.
a b a
Similarly,
Z b Z c Z c
f (x) dx+ f (x) dx ≥ LD1 (f )+LD2 (f ) = LD0 (f ) ≥ LD (f ) ≥ f (x) dx−ε.
a b a
Since ε is arbitrary, it follows that
Z b Z c Z c
f (x) dx + f (x) dx = f (x) dx.
a b a
53
7 The Riemann Integral IA Analysis I
This is different from regular continuity. Regular continuity says that at any
point x, we can find a δ that works for this point. Uniform continuity says that
we can find a δ that works for any x.
It is easy to show that a uniformly continuous function is integrable, since by
uniformly continuity, as long as the mesh of a dissection is sufficiently small, the
difference between the upper sum and the lower sum can be arbitrarily small by
uniform continuity. Thus to prove the above theorem, we just have to show that
continuous functions on a closed bounded interval are uniformly continuous.
Theorem (non-examinable). Let a < b and let f : [a, b] → R be continuous.
Then f is uniformly continuous.
Proof. Suppose that f is not uniformly continuous. Then
Therefore, we can find sequences (xn ), (yn ) such that for every n, we have
1
|xn − yn | ≤ and |f (xn ) − f (yn )| ≥ ε.
n
Then by Bolzano-Weierstrass theorem, we can find a subsequence (xnk ) converg-
ing to some x. Since |xnk −ynk | ≤ n1k , ynk → x as well. But |f (xnk )−f (ynk )| ≥ ε
for every k. So f (xnk ) and f (ynk ) cannot both converge to the same limit. So f
is not continuous at x.
This proof is very similar to the proof that continuous functions are integrable.
In fact, the proof that continuous functions are integrable is just a fuse of this
proof and the (simple) proof that uniformly continuously functions are integrable.
Theorem. Let f : [a, b] → R be monotone. Then f is Riemann integrable.
54
7 The Riemann Integral IA Analysis I
Note that monotone functions need not be “nice”. It can even have in-
finitely many discontinuities. For example, if f : [0, 1] → R maps x to the
1/(first non-zero digit in the binary expansion of x), with f (0) = 0.
ε
Proof. let ε > 0. Let D be a dissection of mesh less than f (b)−f (a) . Then
n
X
UD f − LD f = (xi − xi−1 )(f (xi ) − f (xi−1 ))
i=1
n
ε X
≤ (f (xi ) − f (xi−1 ))
f (b) − f (a) i=1
= ε.
Pictorially, we see that the difference between the upper and lower sums is
total the area of the red rectangles.
y
To calculate the total area, we can stack the red areas together to get something
ε
of width f (b)−f (a) and height f (b) − f (a). So the total area is just ε.
Lemma. Let a < b and let f be a bounded function from [a, b] → R that is
continuous on (a, b). Then f is integrable.
R1
An example where this would apply is 0 sin x1 . It gets nasty near x = 0, but
its “nastiness” is confined to x = 0 only. So as long as its nastiness is sufficiently
contained, it would still be integrable.
The idea of the proof is to integrate from a point x1 very near a up to a
point xn−1 very close to b. Since f is bounded, the regions [a, x1 ] and [xn−1 , b]
are small enough to not cause trouble.
Proof. Let ε > 0. Suppose that |f (x)| ≤ C for every x ∈ [a, b]. Let x0 = a and
ε
pick x1 such that x1 − x0 < 8C . Also choose z between x1 and b such that
ε
b − z < 8C .
55
7 The Riemann Integral IA Analysis I
Proof. Let ε > 0. We need to find an N . The only thing we know is that f is
Riemann integrable, so we use it:
Since f is integrable, there is a dissection D, say u0 < u1 < · · · < um , such
that Z b
ε
UD f − f (x) dx < .
a 2
We also know that f is bounded. Let C be such that |f (x)| ≤ C.
For any n, let D0 be the least common refinement of Dn and D. Then
UD0 f ≤ UD f.
56
7 The Riemann Integral IA Analysis I
Also, the sums UDn f and UD0 f are the same, except that at most m of the
subintervals [xi−1 , xi ] are subdivided in D0 .
For each interval that gets chopped up, the upper sum decreases by at most
b−a
n · 2C. Therefore
b−a
UDn f − UD0 f ≤ 2C · m.
n
Pick n such that 2Cm(b − a)/n < 2ε . Then
ε
UDn f − UD f < .
2
So Z b
UDn f − f (x) dx < ε.
a
4C(b−a)m Rb
This is true whenever n > ε . Since we also have UDn f ≥ a
f (x) dx,
therefore Z b
UDn f → f (x) dx.
a
The proof for lower sums is similar.
For convenience, we define the following:
Notation. If b > a, we define
Z a Z b
f (x) dx = − f (x) dx.
b a
We now prove that the fundamental theorem of calculus, which says that
integration is the reverse of differentiation.
Theorem (Fundamental theorem of calculus, part 1). Let f : [a, b] → R be
continuous, and for x ∈ [a, b], define
Z x
F (x) = f (t) dt.
a
0
Then F is differentiable and F (x) = f (x) for every x.
Proof.
F (x + h) − F (x) 1 x+h
Z
= f (t) dt
h h x
Let ε > 0. Since f is continuous, at x, then there exists δ such that |y − x| < δ
implies |f (y) − f (x)| < ε.
If |h| < δ, then
Z Z
1 x+h 1 x+h
f (t) dt − f (x) = (f (t) − f (x)) dt
h x h x
Z
1 x+h
≤ |f (t) − f (x)| dt
|h| x
ε|h|
≤
|h|
= ε.
57
7 The Riemann Integral IA Analysis I
Proof. Let Z x
g(x) = f 0 (t) dt.
a
Then
d
g 0 (x) = f 0 (x) = (f (x) − f (a)).
dx
Since g 0 (x) − f 0 (x) = 0, g(x) − f (x) must be a constant function by the mean
value theorem. We also know that
So we must have g(x) = f (x) − f (a) for every x, and in particular, for x = b.
Theorem (Fundamental theorem of calculus, part 2). Let f : [a, b] → R be a
differentiable function, and suppose that f 0 is integrable. Then
Z b
f 0 (t) dt = f (b) − f (a).
a
Note that this is a stronger result than the corollary above, since it does not
require that f 0 is continuous.
Proof. Let D be a dissection x0 < x1 < · · · < xn . We want to make use of this
dissection. So write
n
X
f (b) − f (a) = (f (xi ) − f (xi−1 )).
i=1
For each i, there exists ui ∈ (xi−1 , xi ) such that f (xi ) − f (xi−1j ) = (xi −
xi−1 )f 0 (ui ) by the mean value theorem. So
n
X
f (b) − f (a) = (xi − xi−1 )f 0 (ui ).
i=1
We know that f 0 (ui ) is somewhere between sup f 0 (x) and inf f 0 (x)
x∈[xi ,xi−1 ] x∈[xi ,xi−1 ]
by definition. Therefore
LD f 0 ≤ f (b) − f (a) ≤ UD f 0 .
Since f 0 is integrable and D was arbitrary, LD f 0 and UD f 0 can both get arbitrarily
Rb
close to a f 0 (t) dt. So
Z b
f (b) − f (a) = f 0 (t) dt.
a
58
7 The Riemann Integral IA Analysis I
(b − a)2 (2)
f (b) = f (a) + (b − a)f 0 (a) + f (a) + · · ·
2!
Z b
(b − a)n (n) (b − t)n (n+1)
+ f (a) + f (t) dt.
n! a n!
Proof. Induction on n.
When n = 0, the theorem says
Z b
f (b) − f (a) = f 0 (t) dt.
a
59
7 The Riemann Integral IA Analysis I
0
We want something in the form (b − a)f (a), so we take that out and see what
we are left with.
Z b
0 0 0
= f (a) + (b − a)f (a) + b(f (b) − f (a)) − tf (2) (t) dt
a
Rb
Then we note that f 0 (b) − f 0 (a) = a
f (2) (t) dt. So we have
Z b
= f (a) + (b − a)f 0 (a) + (b − t)f (2) (t) dt.
a
Then we can see that the right thing to integrate is (b − t) and continue to obtain
the result.
Theorem (Integration by substitution). Let f : [a, b] → R be continuous. Let
g : [u, v] → R be continuously differentiable, and suppose that g(u) = a, g(v) = b,
and f is defined everywhere on g([u, v]) (and still continuous). Then
Z b Z v
f (x) dx = f (g(t))g 0 (t) dt.
a u
60
7 The Riemann Integral IA Analysis I
when it exists.
Example. Z 1 h i1
x−1/2 dx = 2x−1/2 = 2 − 2ε1/2 → 2.
ε ε
So Z 1
x−1/2 dx = 2,
0
since f is decreasing (the right hand inequality is valid only for n ≥ 2). It follows
that
Z N +1 XN Z N
f (x) dx ≤ f (n) ≤ f (x) dx + f (1)
1 n=1 1
61
7 The Riemann Integral IA Analysis I
P
So
R ∞ if the integral exists, then f (n) is increasing and bounded above by
1
f (x) dx, so converges.
RN PN
If the integral does not exist, then 1 f (x) dx is unbounded. Then n=1 f (n)
is unbounded, hence does not converge.
Rx P∞
Example. Since 1 t12 dt < ∞, it follows that n=1 n12 converges.
62