Analysis on the Real Line
Cam Quilici - Instructor Dr. Andrea Bonito
Fall 2023
Contents
1 The Real Numbers 3
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.1 Axioms of the Real Numbers . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Bounded Intervals . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Supremum and Infimum . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.1 Absolute Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Sequences 8
2.1 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Algebraic Manipulations of Limits . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Comparison Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Limit Infimum and Limit Supremum . . . . . . . . . . . . . . . . . . . . 13
2.5 Subsequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.6 Bolzano-Weirstrass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.7 Cauchy Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.8 Divergent Sequences to ±∞ (Unbounded Case) . . . . . . . . . . . . . . 21
3 Continuous Functions 22
3.1 Sequential Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Algebraic Manipulation of Continuous Limits . . . . . . . . . . . . . . . 25
3.3 Comparison Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 Composition of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5 Infinite Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.6 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.7 Uniform Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.8 Inverses and Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4 Differentiation 42
4.1 Extremum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2 Rules of Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.1 Basic Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.2 Advanced Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Mean Value Theorem and Applications . . . . . . . . . . . . . . . . . . . 49
1
4.4 Limited Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4.1 Local Extremum (Re-Visited) . . . . . . . . . . . . . . . . . . . . . 56
5 Integration 57
5.1 Extension of Integrability . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.2 Properties of Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.3 Antiderivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.4 Properties of Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.5 Change of Variable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.6 Integration by Parts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.7 Taylor Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.8 Weak Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2
1 The Real Numbers
1.1 Introduction
Lemma 1.1.1. The equation x
2
2 =0 has no solution in Q.
Proof. By contradiction, assume there exists some
p
/q Q, p, q N, q 6=0 such
that
µ
p
q
2
2 =0. ()
Without loss of generality, we can assume that the greatest common divisor
between p and q is 1. We rewrite (*) as p
2
=2q
2
which implies that p
2
is even.
This means that p is even as well.
We say that N is well-ordered, but not Q since Q does not have a least element.
Proposition 1.1.2. There is no natural number such that 0 <n <1.
Proof. Left as an exercise.
1.1.1 Axioms of the Real Numbers
Binary operations:
(+): R ×R 7→R x, y R, x +y R,
(·): R ×R 7→R x, y R, x ·y R.
We have axioms for the real numbers as follows:
I Algebraic Axioms
(i) Associativity: For x, y, z R, we have
x +(y +z) =(x +y) +z
x ·(y ·z) =(x ·y)·z
(ii) Commutativity: For x, y R, we have
x +y = y +x
x ·y = y ·x
(iii) Distributivity: For x, y, z R, we have
x ·(y +z) =x ·y +x ·z
3
(iv) Identity:
Addition: There exists some 0 R such that for all x R, x+0 = x.
Multiplication: There exists some 1 R such that for all x R,
x ·1 =x.
(v) Inverses:
Addition: For all x R, there exists some y R such that
x +y = 0 y =x.
Multiplication: For all x R, there exists some y R such that
x ·y = 1 y =
1
x
y = x
1
.
II Ordering
(i) For some x, y, z R, we have
x y = x +z y +z.
(ii) For some x, y R, we have
0 x,0 y = 0 x ·y.
(iii) For some x, y, z R, we have
x y, y z = x z.
(iv) For some x, y R, we have
x y, y x = x = y.
(v) For some x, y R, we have
x 6= y = x y or y x.
III No Hole (not satisfied by Q)
For any non-empty subset X of {x >0}, there exists some a R such that
(1) a x x X , and
(2) For all ε >0(ε 6=0,0 ε), there exists some x
ε
such that x
ε
a ε.
1.1.2 Bounded Intervals
Given a <b, the set X of real numbers is a bounded interval if it is in the form:
Open interval with endpoints a,b
{x X | a < x <b} =(a,b).
Closed intervals
{x X | a x b} =[a,b].
Half-open intervals
{x X | a < x b} =(a,b].
4
1.2 Supremum and Infimum
Definition 1.2.1 (Boundedness). Let X R be non-empty. A number m R
(not necessarily in X ) is a lower bound for X if for all x X , m x.
If a set has a lower bound, it is called bounded below (for instance, Z).
Remark. X either has no lower bounds or infinitely many.
Definition 1.2.2 (Infimum). Let X R be non-empty. A number s R (not
necessarily in X ) is called an infimum of the set X if it is
(1) A lower bound and
(2) For all other lower bounds r
i
of X , s r
i
.
The infimum is hence known as the greatest lower bound." We denote r is
the infimum of X " as
r =inf(X ).
Definition 1.2.3 (Supremum). Let X R be non-empty. A number s R (not
necessarily in X ) is called a supremum of the set X if it is
(1) An upper bound and
(2) For all other upper bounds r
i
of X , s r
i
.
The supremum is hence known as the “least upper bound." We denote r is
the supremum of X " as
r =sup(X ).
Theorem 1.2.4. Let X R be non-empty. X has an upper bound if and
only if X has a finite supremum and the finite supremum is unique.
Proof. If X has a supremum then that supremum is an upper bound for X .
Assume that b is an upper bound of X . Let X ={b+1y | y X }, then X is non
empty and X {x > 0} (there is a 1 so it is surely greater than 0).
So, by Axiom 3, there is a real number a such that
(1) If z X , then a z (i.e., a is a lower bound for X ) and
(2) For every ε >0, there exists a z
ε
X such that
z
ε
a ε.
5
Claim. b
1
=b +1 a is the supremum for X .
There are two things to check:
(1) Is b
1
an upper bound for X ? Well, if y X , then a b +1y and we have
y b +1 a =b
1
.
(2) Is b
1
the least upper bound?
Let M be some upper bound for X . Let us show that b
1
M. By way
of contradiction, assume that M <b
1
=b +1 a. Let ε =b
1
M >0.
By Axiom 3, there exists some z X such that z a ε. We have
z = b +1y, y X , ε =b
1
M =b +1 a M.
So, supposedly z a = b +1 y a b +1 a M with M y (this is
a contradiction since M is assumed to be an upper bound, it cannot be
smaller than y) and therefore b
1
M and b
1
=sup(X ).
Archimedean Axiom (satisfied in R). For every x >0, y 0, there exists n N
such that
n ·x > y.
Proof. By way of contradiction, assume that n ·x y n N
(for some x >
0, y 0). Then let S ={n ·x |n N
} R. Clearly S is bounded above by y and
S is non-empty. This implies that sup(S) exists and n ·x sup(s) n N
. This
all implies that (m +1) ·x sup(S) m N
Lemma 1.2.5. Let X R be non-empty and bounded above. Then for every
ε >0, there exists some x
ε
such that
sup(X )ε <x
ε
sup(X ).
Proof. Since sup(X ) is an upper bound for X , then we know x sup(X ) x
X . This implies the second part of the inequality.
To prove the first part of the inequality, assume by way of contradiction that
there exists some ε
0
> 0 such that for every x X , sup(x) ε
0
x. However,
this is not possible since sup(x) is an upper bound, i.e., subtract anything from
it and it will be less than x.
6
Lemma 1.2.6. Let x, y R with y > x, then there exists some r Q such
that x <r < y.
Proof. Take y x > 0, the Archimidean principle guarantees the existence of
n N such that
n ·(y x) >1 y x >
1
n
.
Recall,
bxc x ≤cxb+1
bnxc
nx
n
<
nx +1
n
.
Let r :=
(bnxc+1)
/n Q. Then we have
x =
nx
n
<
bnxc+1
n
nx +1
n
=x +
1
n
=x +y x
< y.
Corollary. Let p, q Q with p < q. There exists some z R \ Q such that
p <z < q.
Proof. By Lemma,
p
c
<
q
c
p
c
< r
|{z}
Q
<
q
c
p < r ·c
|{z}
R \ Q
<q.
1.2.1 Absolute Value
Definition 1.2.7 (Absolute value). For some x R, the absolute value function
is defined as follows:
|x|:=
(
x , if x <0
x , if x 0.
7
Proposition 1.2.8. Let x,a R. Then we have
(i) |x|=0 x =0.
(ii) Let a >0, then
|x|< a a <x < a
2 Sequences
2.1 Convergence
Definition 2.1.1 (Sequential convergence). A sequence of real numbers (x
n
)
nN
is said to be convergent if there exists x R such that ε > 0 N (ε) N such
that
|x x
n
|ε n N(ε).
Furthermore, we write lim
n→∞
x
n
=x.
Example. Show that lim
n→∞
1 +
1
/(n+1) = 1. We want
|x x
n
|=
¯
¯
¯
¯
1
µ
1 +
1
n +1
¯
¯
¯
¯
=
¯
¯
¯
¯
1
n +1
¯
¯
¯
¯
=
1
n +1
ε ε >0.
Let ε >0, set N (ε) =b
1
/εc=N(ε). This holds for
|x x
n
|=
1
n +1
1
b
1
ε
c+1
1
1
ε
ε.
Lemma 2.1.2. A sequence of real numbers has at most one limit.
Proof. Assume l
1
,l
2
R are two limits. We will show that l
1
= l
2
by showing
that for all ε >0, |l
1
l
2
|ε. Let ε >0.
(1) (x
n
)
nN
converges to l
1
, i.e., for all ε
1
>0 N
1
(ε
1
) such that
|x
n
l
1
|ε
1
n N
1
(ε
1
),
(2) (x
n
)
nN
converges to l
2
, i.e., for all ε
2
>0 N
2
(ε
2
) such that
|x
n
l
2
|ε
2
n N
2
(ε
2
).
8
Let ε >0 and set ε
1
ε
/2,ε
2
=
ε
/2. Then we have
|l
1
l
2
|=|l
1
x
n
+x
n
l
2
|
|l
1
x
n
|+|l
2
x
n
|ε
where |l
1
x
n
| ε
1
n N
1
and |l
2
x
n
| ε
2
n N
2
n max{N
1
, N
2
}. Alto-
gether, this implies that l
1
=l
2
.
Example of Divergent Sequence. Take x
n
= n ·sin(n ·
π
/2). By way of contra-
diction, assume that (x
n
)
nN
is converging to some x R. Say
x
p
n
=x
4n+1
=(4n +1) ·sin((4n +1) ·
π
2
)
| {z }
=1
=4n +1
x
q
n
=x
4n+5
=(4n +5) ·sin((4n +5) ·
π
2
)
| {z }
=1
=4n +5
which implies that |x
p
n
x
x
n
| = 4. From the contradiction assumption, for
ε =
1
/2 there exists N (
1
/2) such that
|x
n
x|
1
2
n N(
1
2
).
Take n such that 4n +1 N (
1
/2). Then we have
|x
p
n
x|
1
2
|x
q
n
x|
1
2
4 =|x
p
n
x
q
n
|=|x
p
n
x +x x
q
n
|
|x
p
n
x|+|x
q
n
x|
1
2
+
1
2
1
but this is a contradiction, so (x
n
)
nN
must be divergent.
Lemma 2.1.3. Every convergent sequence is bounded.
Proof. Let (x
n
)
nN
be a convergent sequence, that is, there exists some x R
such that for all ε >0, there exists some N(ε) N and
|x x
n
|ε n N(ε).
9
In particular, take for ε =1, there is N (1) such that
|x
n
x|1 n N(1).
Recall the reverse triangle inequality,
|x
n
||x||x x
n
| 1 n N(1)
= |x
n
|1 +|x| n N (1)
Let M =max{|x
0
|,|x
1
|,. .. ,|x
N(1)1
|,1 +|x|}. Then certainly
|x
n
|M n N(1)
|x
n
|M n N(1)
= | x
n
|M n N.
Recall. A sequence (x
n
)
nN
convergent = M N such that |x
n
|M n N.
Definition 2.1.4 (Sequential monotonicity). A sequence (x
n
)
nN
is said to be
Increasing if x
m
x
n
when m n,
Decreasing if x
m
x
n
when m n,
Monotone if (x
n
)
nN
is increasing or decreasing.
Lemma 2.1.5. Any increasing sequence that is bounded above is conver-
gent.
Proof. Let X = {x
n
| n N}. Then x
n
M,n N = X bounded above =
x = sup(X ).
We want to show that (x
n
)
nN
converges to x =sup(X ). This means (by defini-
tion) that for every ε >0, there exists N (ε) N such that
|x
n
x|ε n N(ε).
Note that x =sup(X ) implies two important things:
(i) x
n
x n N,
(ii) ε >0, x
ε
X such that xx
ε
ε by the characterization of supremum.
Note, we can equivalently replace x
ε
with x
n
ε
N to obtain ε >0, x
n
ε
X such that x x
n
ε
ε.
10
So, let ε >0, compute
|x
n
x|= x x
n
x x
n
ε
ε
with x
n
ε
x
n
and n n
ε
((x
n
)
nN
increasing) and for all ε > 0, take N (ε) =
n
ε
.
Corollary. A decreasing sequence (x
n
)
nN
that is bounded below is conver-
gent.
Remark. There are two cases:
(i) (x
n
)
nN
increasing, bounded above = lim
n→∞
x
n
=sup{x
n
|n N},
(ii) (x
n
)
nN
decreasing, bounded below = lim
n→∞
x
n
=inf{x
n
|n N}.
Example. Take x
n
=2
n
decreasing, bounded below. Then
lim
n→∞
x
n
=inf{2
n
|n N} =0.
Corollary. Let (x
n
)
nN
be a monotone sequence. Then (x
n
)
nN
is convergent
if and only if (x
n
)
nN
is bounded.
2.2 Algebraic Manipulations of Limits
Lemma 2.2.1. Let (x
n
)
nN
,(y
n
)
nN
be convergent sequences. Set lim
n→∞
x
n
=
x,lim
n→∞
y
n
= y. Three things are true:
(1) The sequence (x
n
+y
n
)
nN
is convergent and li m
n→∞
(x
n
+y
n
) =x+y,
(2) The sequence (x
n
·y
n
)
nN
is convergent and lim
n→∞
(x
n
·y
n
) =x ·y,
(3) Assume x
n
6=0, y
n
6=0, then the sequence (
y
n
/x
n
)
nN
is convergent and
lim
n→∞
(
y
n
/x
n
) =
y
/x.
Proof. of (2). We want to show that ε >0 N (ε) N such that
|x
n
y
n
x y|ε n N (ε).
Note that
|x
n
y
n
x y|=|(x
n
x)y
n
+x(y
n
y)|
|x
n
x||y
n
|+|x||y
n
y|
by the triangle inequality.
11
Case 1: x 6= 0. Because (y
n
)
nN
is convergent, there exists N
1
N such
that |y
n
y|
ε
/2·|x| n N
1
. Also, (y
n
)
nN
is bounded, i.e., |y
n
|M n
N M 1.
Because (x
n
)
nN
is convergent, there exists some N
2
N such that
|x x
n
|
ε
2M
n N
2
.
Therefore, for all ε >0, we set N
ε
=max{N
1
, N
2
} and we have that
|x
n
y
n
x y|
ε
2
+
ε
2
ε n N
ε
which ultimately shows that lim
n→∞
x
n
·y
n
=x ·y.
Case 2: x = 0. Obviously it is true that
|x
n
y
n
x y|=|x
n
y
n
||x
n
||y
n
|.
Then, consider that
*
(y
n
)
nN
convergent implies |y
n
|M , and
*
(x
n
)
nN
convergent to 0 implies ε > 0 N
1
N such that |x
n
|
ε
/2M n N
1
. Again, we have
x
n
y
n
x y|
ε
2M
·M
ε
2
ε n N
1
which implies that lim
n→∞
x
n
y
n
=0.
2.3 Comparison Results
Lemma 2.3.1. Let (x
n
)
nN
and (y
n
)
nN
be two convergent sequences such
that there exists N
0
N such that
x
n
y
n
n N
0
.
Then
lim
n→∞
x
n
lim
n→∞
y
n
.
Proof. Let x := lim
n→∞
x
n
, y :=lim
n→∞
. For the sake of contradiction, assume
that y < x. Because lim
n→∞
x
n
= x and lim
n→∞
y
n
= y, we have that for ε =
(xy)
/4, there exists some N N such that
|x
n
x|ε and |y
n
y|ε n N.
12
Note that
|x
n
x|ε = x x
n
+ε
|y
n
y|ε = y y
n
ε.
This means that for n max{N
0
, N}, we have
y
n
y +ε = y +
x y
4
< y +
x y
2
| {z }
(x+y)/2
=x
x y
2
x
x y
4
=x ε <x
n
which implies that y
n
<x
n
which is a contradiction.
Theorem 2.3.2 (Squeeze Theorem). Let (x
n
)
nN
, (y
n
)
nN
, (z
n
)
nN
be se-
quences such that
Sequences (x
n
)
nN
and (z
n
)
nN
are both converging to some limit L,
and
There exists some N
0
N such that
x
n
y
n
z
n
n N
0
.
Then (y
n
)
nN
is convergent and lim
n→∞
=L.
Proof. Since lim
n→∞
x
n
=L and lim
n→∞
z
n
=L, we have that ε >0 N (ε) N
such that
|x
n
L|ε and |z
n
L|ε n N(ε).
Then, consider N max{N (ε), N
0
} and we have
L ε x
n
y
n
z
n
L +ε
i.e., Lε y
n
L+ε or |Ly
n
|ε n max{N(ε),N
0
} = lim
n→∞
y
n
=L.
2.4 Limit Infimum and Limit Supremum
Let (x
n
)
nN
be a bounded sequence. Construct a new sequence y
n
={x
k
: k
n}. We want to show two things about this sequence:
13
(1) (y
n
)
nN
is decreasing:
Consider n <m, then we have y
n
=sup{x
k
: k n}
|
{z }
A
and y
m
=sup{x
k
: k m}
| {z }
B
.
It is clear that B A = sup(B ) sup(A) (by an exercise) which alto-
gether implies that y
m
y
n
.
(2) (y
n
)
nN
is bounded below:
Since (x
n
)
nN
is bounded below, so is (y
n
)
nN
.
Definition 2.4.1 (limsup and liminf). Let (x
n
)
nN
be a sequence that is bounded
above. Then we define the limit superior as
limsup
n→∞
x
n
= lim
n→∞
sup{x
k
: k n}
| {z }
y
n
.
Similarly, when (x
n
)
nN
is bounded above, we define the limit inferior as
liminf
n→∞
x
n
= lim
n→∞
inf{x
k
: k n}
| {z }
y
n
.
Example. Consider the sequence x
n
=(1)
n
. Then we have
limsup
n→∞
x
n
= lim
n→∞
sup{x
k
: k n} =1
liminf
n→∞
x
n
= lim
n→∞
inf{x
k
: k n} =1.
Since limsup
n→∞
x
n
6=liminf
n→∞
x
n
, (x
n
)
nN
diverges.
Theorem 2.4.2. Let (x
n
)
nN
be a bounded sequence. The sequence (x
n
)
nN
is convergent if and only if
limsup
n→∞
x
n
=liminf
n→∞
x
n
.
Proof. Assume that (x
n
)
nN
is convergent and let x := lim
x→∞
x
n
. Further, de-
note
z
n
=inf{x
k
: k n}
z = lim
n→∞
z
n
=liminf
n→∞
x
n
.
14
(we want to show that x = z.) Since x
n
converges to x, we know that for every
ε >0, there exists some N
1
(ε) N such that
|x x
n
|
ε
2
n N
1
(ε)
and since z
n
converges to z we similarly have that there exists some N
2
(ε) N
such that
|z z
n
|
ε
4
n N
2
(ε).
Recall that z
n
= inf{x
k
: k n}. Then, thanks to the characterization of in-
fimums, we know that for every ε > 0 (pick our ε =
ε
/4), there exists some
x
m
ε
{x
n
: k n} such that
|x
m
ε
z
n
|=x
m
ε
z
n
ε
4
since m
ε
n and z
n
x
m
ε
.
Now, compute
|x z|=|x x
m
ε
+x
m
ε
z
n
+z
n
z|
|x x
m
ε
|
| {z }
ε
/2
+|x
m
ε
z
n
|
| {z }
ε
/4
+|z
n
z|
| {z }
ε
/4
ε m n N
1
(ε).
So we proved that ε >0,
|x z|ε
= x z =0
= x = z.
From a similar argument, lim
n→∞
x
n
= lim sup
n→∞
x
n
= y, i.e., x = y = y =
z.
The other direction is much simpler. Assume y =limsup
n→∞
x
n
=liminf
n→∞
x
n
=
z. Then, recall
y y
n
=sup{x
k
: k n}
z z
n
=inf{x
k
: k n}.
So, by our assumption we have z
n
x
n
y
n
which by the squeeze theorem
implies that z = y. Therefore, x
n
y = z i.e., lim
n→∞
x
n
= y = z.
2.5 Subsequences
15
Definition 2.5.1 (Subsequences). A subsequence (x
n
)
nN
is a sequence (y
n
)
nN
such that for all k N, there exists n
1
,n
2
,. .. ,n
k
N with n
1
<n
2
<···<n
k
with
y
k
=x
n
k
.
Lemma 2.5.2. Any subsequence (x
n
k
) of a convergent sequence (x
k
)
kN
is
convergent to the same limit.
Proof. Left as an exercise.
Lemma 2.5.3. Let (n
k
)
k=1
be a sequence of increasing natural numbers.
Then
n
k
k.
Proof. By induction (left as an exercise).
Proof. (Of first Lemma). For every ε >0, there exists some N(ε) N such that
|x x
k
|ε k N(ε)
(because x
k
x). From previous Lemma (above), n
k
k N (ε) so we have
|x x
n
k
|{z}
=y
k
|ε n
k
N (ε).
Lemma 2.5.4. Any subsequence of a bounded sequence is bounded.
Proof. Left as an exercise.
2.6 Bolzano-Weirstrass
Definition 2.6.1 (Peak points). A peak point of a sequence (x
n
)
nN
is a term
x
p
such that x
p
>q
p
means q > p.
Lemma 2.6.2 (Monotone Subsequence). Every sequence has a monotone
subsequence.
16
Proof. We argue on whether the sequence has either (A) no peak point, (B)
finitely many peak points, or (C) infinitely many peak points.
(A) We construct (x
n
k
)
kN
as follows:
n
1
=1, x
n
1
=x
1
is not a peak point (no peak points)
= q >1 such that x
q
>x
1
>x
n
1
.
n
2
=q, x
n
2
=x
q
.
.
.
x
n
1
x
n
2
···montonically increasing
(B) If x
p
is the final peak point (p is the largest index where x
p
is a peak
point), then take n
1
=p +1, x
n
1
=x
p+1
. Then proceed as before.
(C) In this case
p
1
<p
2
<···<p
n
<··· .
By definition, x
p
i
> x
p
j
for j > i . So we have a decreasing subsequence
taking x
n
k
=x
p
k
.
Theorem 2.6.3 (Bolzano-Weirstrass). Every bounded sequence has a con-
vergent subsequence.
0
M M
I
1
I
2
a
n
1
a
n
2
Proof. We have that (x
n
)
nN
implies that (x
n
k
)
kN
is monotone. Further, (x
n
)
nN
bounded implies that (x
n
k
)
kN
is bounded. So we have a montone and bounded
subsequence which is thus convergent.
Lemma 2.6.4. Let (x
n
)
nN
be a bounded sequence. There is a convergent
subsequence x
n
k
such that
lim
k→∞
x
n
k
=limsup
n→∞
x
n
.
17
Proof. We proceed by induction. Let y = lim
n→∞
y
n
, y
n
= sup{x
k
: k n}
such that y = lim sup
n→∞
y
n
.
P(k): There are n
1
<n
2
<···<n
k
and x
n
1
, x
n
2
,. .. , x
n
k
such that
y
1
i
|{z}
y
<x
n
i
< y +
1
i
|{z}
y
1 i k
which implies that x
n
i
y as well by squeeze theorem.
P(1): There is n
1
N and x
n
1
such that y 1 <x
n
1
< y +1.
We have y
n
y = N N such that |y
n
y|
1
/4 n N. Further we
have y
n
= sup{x
n
: k n} = x
m
,m n such that y
n
x
m
1
/4 by the
characterization of suprema. Then compute
|y x
n
|=|y y
n
+y
n
x
m
|
|y y
n
|+|y
n
x
m
|
1
4
+
1
4
=
1
2
so y
1
/2 x
m
y +
1
/2 = y 1 <x
m
< y +1.
Next, assume P (k) is true: There is n
1
<n
2
<···<n
k
and x
n
1
, x
n
2
,. .. , x
n
k
such
that
y
1
i
<x
n
i
< y +
1
i
1 i k.
Since y
n
y, N N such that |y y
n
|
1
/(4(k+1)), n max{N ,n
k+1
} > N
k
.
Further, we have y
n
= sup{x
k
: k n} which implies there exists some m N
with m n > n
k
such that
y
n
x
m
<
1
4(k +1)
.
Therefore, we have
|y y
m
||y y
n
|+|y
n
x
m
|
1
4(k +1)
+
1
4(k +1)
=
1
2(k +1)
<
1
k +1
= y
1
k +1
<x
m
< y +
1
k +1
,n
k+1
=m >n
k
.
Therefore, P(k +1) is true. Thus, we have shown that P (k) is true for all k N
and x
n
k
y by the squeeze theorem.
18
Remark. Every bounded sequence (x
n
)
nN
has a convergent subsequence
(x
n
k
)
kN
such that
lim
k→∞
x
n
k
=liminf
n→∞
x
n
.
(Proof is the same).
Result (forgot to mention last class). If (x
n
)
nN
is bounded, then every conver-
gent subsequence (x
n
k
)
kN
satisfies
liminf
n→∞
x
n
lim
k→∞
x
n
k
limsup
n→∞
x
n
.
Example. Take x
n
=(1)
n
. We want to find the limsup.
Show that (x
n
)
nN
is bounded, i.e., x
n
1 = limsup
n→∞
1. Then, con-
sider the subsequence (x
2k
) = (1)
kN
. Then show that (x
2k
)
k→∞
1 = 1
limsup
n→∞
x
n
.
Since we have shown that limsup
n→∞
x
n
1 lim sup
n→∞
x
n
, we know that
limsup
n→∞
x
n
=1.
2.7 Cauchy Sequences
Definition 2.7.1 (Cauchy sequences). A sequence (x
n
)
nN
is Cauchy if ε >
0,N N with
|x
m
x
n
|ε m, n N.
Example. Take x
n
=
1
/2(x
n1
+x
n2
), x
0
=0, x
1
=1. Then we compute
x
n
x
n1
=
1
2
(x
n1
+x
n2
) x
n1
=
1
2
(x
n2
x
n1
)
x
2
x
1
=
1
2
= x
2
=
1
2
+1 =
1
2
x
3
x
2
=
1
2
(1
1
2
) =
µ
1
2
2
.
Then, by induction we can show that
x
n
x
n1
=
µ
1
2
n1
.
19
Now take m n and compute
x
m
x
n
=x
m
x
m1
+x
m1
x
m2
±···±x
n+1
x
n
|x
m
x
n
|
µ
1
2
m1
+
µ
1
2
m2
+···+
µ
1
2
n
=
µ
1
2
n
µ
1 +
1
2
+
µ
1
2
2
+···+
µ
1
2
mn1
=
1
¡
1
2
¢
m2
1
1
2
2 (by geometric sum)
2
n+1
.
Given ε >0, if we want 2
n+1
ε, compute
2
n+1
ε
n +1 ln(ε)
n ln(ε) +1.
Therefore, take N = bln(ε)c+2, n ln(ε) +1. So, for m,n N = |x
m
x
n
|
ε = x
n
Cauchy. In fact, lim
n→∞
x
n
=
2
/3 (details left to reader).
Warning: Checking the difference between two subsequent terms of the se-
quence is not enough to ensure that the sequence is Cauchy.
For example, consider x
n
=
P
n
k=1
1
/k (which is not Cauchy).
Lemma 2.7.2. A sequence (x
n
)
nN
is convergent if and only if (x
n
)
nN
is
Cauchy.
Proof. First, assume that x
n
x. Therefore, ε >0, N N such that
|x
n
x|
ε
2
n N.
Therefore, for ε >0 and m,n N ,
|x x
m
|
ε
2
and |x x
n
|
ε
2
.
This implies that
|x
m
x
n
|=|x x
m
+x x
n
|
|x x
n
|+|x x
m
|
ε
2
+
ε
2
=ε
= x
n
Cauchy.
20
For the other direction, assume (x
n
)
nN
is Cauchy. First, we must prove some
claims that will lead us to this proof.
(1) Claim. If (x
n
)
nN
is Cauchy, then (x
n
)
nN
bounded.
Proof. This proof is omitted because it is almost identical to the proof as
convergent implies bounded.
(2) Claim. If (x
n
)
nN
has a convergent subsequence (x
n
k
)
kN
, then (x
n
)
nN
is convergent (to the same limit).
Proof. To see this, assume there exists some (x
n
k
) convergent subsequence
and
(i) We know x
n
is Cauchy, i,eε >0, N
1
N such that
|x
m
x
n
|
ε
2
m, n N
1
.
(ii) There exists an x N such that ε >0, N
2
N such that
|x x
n
k
|
ε
2
k N
2
.
Now, we will show that the entire sequence (x
n
) converges to x.
For every ε >0, set N =max N
1
, N
2
. Then compute
|x x
k
||x x
n
k
|+|x
n
k
x
k
|
ε
2
+
ε
2
k N .
So, we have
(x
n
) Cauchy
(1)
=(x
n
) bounded
(B.W.)
===(x
n
) has a convergent subsequence
(2)
=(x
n
) convergent.
2.8 Divergent Sequences to ±∞ (Unbounded Case)
Definition 2.8.1 (Divergence to ±∞). We say that a sequence (x
n
)
nN
is di-
verging to +∞ if M R, N N such that
x
n
M n M.
Similarly, we say that a sequence is diverging to −∞ if M R, N N such
that
x
n
M n N.
21
Lemma 2.8.2. If (x
n
)
nN
is converging to ±∞, then every subsequence is
convering to ±∞ (respectively).
3 Continuous Functions
Definition 3.0.1 (Functional limits). Let a < b, x
0
(a,b), f : (a,b) \ {x
0
} 7→ R.
We say that f has a limit l at x
0
if ε >0 δ >0 such that, for x (a,b)
0 <|x x
0
|δ = |f (x) l|ε.
If f has a limit l, we write
lim
xx
0
f (x) =l.
a δ
a
a +δ
L ε
L
L +ε
x
y
Proposition 3.0.2. Let x
0
(a,b), f : (a,b)\ {x
0
} 7→R. If f has a limit at x
0
,
then the limit must be unique.
Proof. For the sake of contradiction, assume that l
1
,l
2
are two limits of f at x
0
and further assume that l
1
<l
2
. Take ε =(l
2
l
1
)/4. Then we have
l
1
a limit means δ
1
>0 such that
0 <|x x
0
|δ
1
= |f (x) l
1
|ε.
l
2
a limit means δ
2
>0 such that
0 <|x x
0
|δ
2
= |f (x) l
2
|ε.
22
Then, compute for 0 <|x x
0
|min{δ
1
,δ
2
}
l
2
l
1
|l
2
f (x)|+|l
1
f (x)|
ε +ε =2ε
=
l
2
l
1
2
<l
2
l
1
.
Of course, this is a contradiction so l
1
=l
2
.
Example. Take f (x) = x
2
. We wish to show that lim
x2
f (x) =4. To show, we
have ε >0 δ >0 such that
0 <|x 2|δ = |x
2
4|ε.
Then, compute
|x
2
4|=|x 2|
|{z}
δ
|x +2|
δ(|x|+2)
δ(4 +δ)
= | x|2 +δ
(since |x|= |x 2 +2| |x 2|+2). Now we want δ (4 +δ) ε. Further, restrict
δ 1 = δ
2
δ and δ(4 +δ) 5δ. We must now show that 5δ ε.
So, ε >0, set δ =min{1,
ε
/5}. Then
0 <|x 2|δ = |x
2
4|5δ ε.
So indeed lim
x2
f (x) =4.
Example. Take f (x) =(x1)/(x1), x 6=1. This function is not defined at x =1,
so take f : (2,2) \ {1} 7→ R. However, we claim still that lim
x→∞
f (x) = 1. We
show ε >0, set δ =1. Then we have
0 <|x 1|δ = |f (x) 1|=|1 1|=0 ε.
3.1 Sequential Limits
Proposition 3.1.1. Let x
0
(a,b), f : (a,b)\{x
0
} 7→R. Then f has the limit l
at x
0
if and only if for every sequence (z
n
)
nN
, z
n
(a,b)\{x
0
} with z
n
x
0
,
then the sequence (f (z
n
))
nN
has
f (z
n
) l.
23
Proof. First, suppose lim
xx
0
f (x) = l and (z
n
)
nN
, z
n
(a,b) \ {x
0
} with z
n
x
0
. Then, we know the following
(1) lim
xx
0
f (x) =l = ε > 0 δ > 0 such that 0 < |x x
0
| δ = |f (x)
l|ε and
(2) z
n
x
0
=
e
ε
e
N N such that |z
n
x
0
|
e
ε n
e
N.
Then, let ε >0 be given. Set
e
ε = δ so that n N = 0 <|z
n
x
0
| δ and then
together with (1) this implies |f (z
n
)l|ε. Of course, this means lim
n→∞
f (z
n
) =
l by definition.
For the other direction, assume by contradiction that lim
xx
0
f (x) 6=l. By def-
inition, this implies that ε
0
0 such that δ > 0 and 0 < |x x
0
| δ =
|f (x) l | > ε
0
. Now, let n N and z
n
(x
0
1
/n, x
0
+
1
/n) \{x
0
}. By the Squeeze
Theorem (Theorem 2.3.2)
x
0
1
n
z
n
x
0
+
1
n
= z
n
x
0
.
By assumption, lim
n→∞
f (z
n
) =l, i.e., ε >0 n N such that n N , |f (z
n
)
l|ε. Taking δ =
1
/n, we obtain a contradiction. Therefore, our contradiction
assumption must be false and lim
xx
0
f (x) =l.
Example. Take f (x) =
x
/|x|, x 6= 0. How do we show that this does not have a
limit at x = 0? We could use the epsilon-delta proof technique, but this could
be long and tedious. Instead, consider
Let z
n
=
1
n
0
f (z
n
) =1 1
Let
f
z
n
=
1
n
0
f (
f
z
n
) =1 1
= f (x) does not have a limit at x
0
.
Corollary. Let x
0
(a,b) \ {x
0
}, f : (a,b) 7→ R. Assume that every sequence
(z
n
)
nN
, z
n
(a,b) \ {x
0
}, z
n
x
0
, one has (f (z
n
))
nN
is convergent. Then f
has a limit at x
0
.
Proof. We will show that for (v
n
)
nN
, (w
n
)
nN
two sequences with v
n
, w
n
(a,b) \ {z
0
} and v
n
x
0
, w
n
x
0
, one has lim
n→∞
f (v
n
) =lim
n→∞
f (w
n
).
For the sake of contradiction, assume l
1
=limn f (v
n
) 6=lim
n→∞
f (w
n
) =
l
2
. Set
y
n
=
(
v
n
n even
w
n
n odd.
24
Then we have (1) y
n
(a,b) \ {x
0
} and (2) y
n
x
0
together imply ( f (y
n
))
nN
convergent. However,
f (y
n
) =
(
f (v
n
) n even
f (w
n
) n odd
cannot be convergent since f (y
2n
) l
2
and f (y
2n+1
) l
2
which contradicts
our assumption that l
1
6= l
2
. Therefore, our assumption must be false and
l
1
=l
2
. Combining this with our previous result (Proposition 3.1.1) we get that
lim
xx
0
f (x) =l
1
=l
2
.
Lemma 3.1.2 (Cauchy Criteria for Functions). Let x
0
(a,b)\{x
0
}, f : (a,b)\
{x
0
} 7→R. Then f has a limit at x
0
if and only if for every ε > 0, there exists
δ >0 such that for every x
1
, x
2
satisfying
0 <|x
1
x
0
|δ 0 <|x
2
x
1
|δ,
one has
|f (x
1
) f (x
2
)|ε.
Proof. First, assume that lim
xx
0
f (x) =l. This implies that ε >0 δ >0 such
that |0 < |x x
0
| δ = |f (x) l|
ε
/2. Let ε > 0 be given, then there exists
some δ >0 such that 0 <|x
1
x
2
| δ and 0 <|x
2
x
1
| δ = |f (x
i
) l |
δ
/2
for i =1,2. Therefore,
|f (x
1
) f (x
2
)||f (x
1
) l|+|f (x
2
) l|
ε
2
+
ε
2
=ε.
For the other direction, we will show that for every sequence (z
n
)
nN
, z
n
(a,b) \ {x
0
}, (f (z
n
))
nN
is convergent (Corollary implies f has a limit at x
0
).
Recall the fact that (f (z
n
))
nN
convergent (f (z
n
))
nN
Cauchy. Thus, we
only need to show that ε >0 n N such that m,n N , |f (z
n
) f (z
m
)|ε.
Since z
n
x
0
by assumption, we also know that for every δ > 0, N N such
that n N, 0 <|z
n
x
0
|δ.
Putting everything together, we get that ε > 0, N N and 0 < |x
0
z
n
|
δ n N which in turn implies that |f (z
n
) f (z
m
)| ε m,n N. So, we
know that (f (z
n
))
nN
is Cauchy. Therefore, (f (z
n
))
nN
is convergent.
3.2 Algebraic Manipulation of Continuous Limits
The theme of this subsection is: everything we know about sequences holds
for limits of continuous functions."
25
Lemma 3.2.1. Let f , g : (a,b)\{x
0
} 7→R with lim
xx
0
f (x) =l
1
and lim
xx
0
g (x) =
l
2
. Then
(1) f +g has a limit at x
0
and lim
xx
0
(f +g ) =l
1
+l
2
.
(2) For λ R, λ · f has a limit at x
0
and lim
xx
0
(λ · f ) =λ·l
1
.
(3) f ·g has a limit at x
0
and lim
xx
0
(f ·g ) =l
1
·l
2
.
(4) Assume f (x) 6=0,x (a,b) \ {x
0
} and l
1
6=0. Then
g
/f has a limit at x
0
and lim
xx
0
(
g
/f ) =
l
2
/l
1
.
Proof. (1) Let (z
n
)
nN
, z
n
(a,b) \ {x
0
}, z
n
x
0
. We know that
lim
xx
0
f (x) =l
1
= f (z
n
) l
1
,
lim
xx
0
g (x) =l
2
= g(z
n
) l
2
.
Then, using the properties of sequences,
lim
n→∞
(f (z
n
) +g (z
n
)) =l
1
+l
2
which implies that lim
xx
0
(f +g )(x) =l
1
+l
2
.
3.3 Comparison Results
Lemma 3.3.1. Let f , g : (a,b) \ {x
0
} 7→R. Assume further that
(1) lim
xx
0
f (x) =l
1
and lim
xx
0
g (x) =l
2
and
(2) α >0 such that 0 <|x x
0
|α = f (x) g (x).
Then l
1
l
2
.
Proof. Let (z
n
)
nN
, z
n
(a,b)\{x
0
}, z
n
x
0
. This tells us that N N such that
n N, 0 <|z
n
x
0
|α
(2)
= f (z
n
) g (z
n
) n N.
Then, using the properties of sequences, we get that
l
1
= lim
n→∞
f (z
n
) lim
n→∞
g (z
n
) =l
2
.
Lemma 3.3.2. Let f : (a,b) \ {x
0
} 7→ R such that lim
xx
0
f (x) = l . Then
lim
xx
0
|f (x)|=|l|.
26
Proof. By the reverse triangle inequality, we have 0 ||f ||l || |f l|. Then,
let
f (x) =
(
1 x Q
1 x R \ Q.
Then |f (x)|=1 = lim
x0
|f |(x) =1. But δ >0 in (0,δ) there is always
x
1
Q (0,δ)
x
2
R \ Q (0,δ).
We know further that |f (x
1
) f (x
2
)| = 2 which implies that f does not have
a limit at 0 since x
1
and x
2
do not satisfy the criteria of a Cauchy sequence
(Lemma 3.1.2).
Theorem 3.3.3 (Squeeze Theorem). Let f ,g,h : (a,b) \ {x
0
} 7→ R. Further,
assume
(1) lim
xx
0
f (x) =l =lim
xx
0
g (x) and
(2) α >0 such that 0 <|x x
0
|α.
Then lim
xx
0
h(x) =l.
Proof. Let (z
n
)
nN
, z
n
(a,b) \ {x
0
}, z
n
x
0
. This means N N such that
n N, 0 <|z
n
x
0
|α. By (2) of Lemma 3.3.1, we have that
f (z
n
) h(z
n
) g (z
n
) n N.
Therefore, we have lim
n→∞
h(z
n
) =l and because (z
n
)
nN
is arbitrary, lim
xx
0
h(x) =
l.
Example. Take f : (
π
/2,
π
/2) \ {0} 7→ R defined by f (x) =
sin x
/x, x > 0. How can
we show that lim
x0
f (x) = 1? We could show this rigorously with the defini-
tion of limits, however this would be tedious. Instead, consider that we know
sin x x. In particular,
|1 cos x|=2
¯
¯
¯
sin
2
x
2
¯
¯
¯
x
2
2
.
This implies that lim
x0
cos x = 1. On the other hand,
sin x
2
π ·
x
2π
tan x
2
and for x 6=0, we obtain
1
x
sin x
1.
Since cos x
x0
1 and 1
10
1, we have by the squeeze theorem for functions
that lim
x0
sin x
/x =1.
27
3.4 Composition of Functions
Lemma 3.4.1 (Composition of Functions). Let f : (a,b)\{x
0
} 7→(c,d) such
that lim
xx
0
f (x) = y
0
. Let g : (c,d)\{y
0
} 7→R and assume that lim
yy
0
g (y) =
l. Then g f has a limit at x
0
and lim
xx
0
(g f )(x) =l.
Note: we must assume α > 0 such that 0 <|x x
0
|α = f (x) = y
0
).
Proof. We know two things by assumption:
(1) lim
xx
0
f (x) = y
0
means
ε
1
>0 δ
1
>0 such that 0 <|x x
0
|δ
1
= |f (x) y
0
|ε
1
.
(2) lim
yy
0
g (y) =l means
ε
2
>0 δ
2
>0 such that 0 <|y y
0
|δ
2
= | g (y) l|ε
2
.
Here, we further assume that α >0 such that 0 <|x x
0
|α = f (x) 6=
y
0
. Altogether this implies that
0 |x x
0
|min{δ
1
,α} = 0 <|f (x) y
0
|ε
1
.
So we want to show that ε >0 δ >0 such that
0 <|x x
0
|δ = |g ( f (x)) l| ε.
Let ε > 0 be given. Set ε
2
= ε = δ
2
, set ε
1
= δ
2
= δ
1
and then set δ =
min{δ
1
,α}. Thus,
0 <|x x
0
|δ = 0 <| f (x)
|{z}
=y
y
0
|ε
1
=δ
2
= | g ( f (x)) l|ε.
3.5 Infinite Limits
Definition 3.5.1 (Infinite functional limits). We say that f : (a,b) \ {x
0
} 7→ R
tends to +∞ when x tends to x
0
if M R δ >0 such that
0 <|x x
0
|δ = f (x) M.
We write lim
xx
0
f (x) =+∞.
Similarly, we say that f tends to −∞ when x tends to x
0
if M R δ > 0
such that
0 <|x x
0
|δ = f (x) M.
We write lim
xx
0
f (x) =−∞.
28
Definition 3.5.2 (Infinite functional limits). We say that f : (a,+∞) 7→R has a
limit l when x tends to +∞ if ε >0 D R such that
x D = |f (x) l|ε.
We write lim
x→+∞
f (x) =l.
Similarly, we say that f : (−∞, b) 7→ R has a limit l when x tends to −∞ if
ε >0 D R such that
x D = |f (x) l|ε.
We write lim
x→−∞
f (x) =l.
Definition 3.5.3 (Combined infinite functional limits). We say that f : (a,+∞) 7→
R tends to +∞ when x +∞ if M R D R such that
x D = f (x) M .
We write lim
x→+∞
f (x) =+∞.
Similarly, we say that f : (−∞,b) 7→ R tends to −∞ when x −∞ if M
R D R such that
x D = f (x) M .
We write lim
x→−∞
f (x) =−∞.
3.6 Continuity
Definition 3.6.1 (Functional continuity). We say that f : (a,b) 7→ R is contin-
uous at x
0
(a,b) if
lim
xx
0
f (x) = f (x
0
).
This is equivalent to:
(i) ε >0 δ(ε, x
0
) >0 such that |x x
0
|δ = |f (x) f (x
0
)|ε.
Proof. Left as an exercise.
(ii) ε > 0 δ(ε, x
0
) > 0 such that |x
1
x
0
| δ and |x
2
x
0
| δ = |f (x
1
)
f (x
2
)|ε.
Proof. Left as an exercise.
29
Lemma 3.6.2. Let x
0
(a,b) and f : (a, b) 7→R. Then f is continuous at x
0
if and only if for every sequence (z
n
)
nN
, z
n
(a,b), z
n
x
0
,
f (z
n
) f (x
0
).
Proof. First, assume that f is continuous at x
0
. In other words, lim
xx
0
f (x) =
f (x
0
). Let the sequence (z
n
)
nN
, z
n
(a,b), z
n
x
0
. Let ε > 0 be given. Then
N N such that n N (z
n
6=x
0
)
,
|f (z
n
) f (x
0
)|ε
= f (z
n
) f (x
0
).
For the other direction, assume that every sequence (z
n
)
nN
, z
n
(a,b), z
n
x
0
= f (z
n
) f (x
0
). In particular, consider the sequences (z
n
)
nN
, z
n
(a,b) \ {x
0
}, z
n
x
0
= f (z
n
) f (x
0
) = lim
xx
0
f (x) = lim
n→∞
f (z
n
) =
f (x
0
).
Example. Take f : (0, +∞) 7→ R be defined by f (x) =
p
x. We claim that f is
continuous at every x
0
(0,+∞). Set x (0,+∞and compute
|f (x) f (x
0
)|=|
p
x
p
x
0
|=
|x x
0
|
p
x
|{z}
>0
+
p
x
0
|x x
0
|
p
x
0
.
Now, ε >0 set δ =ε·
p
x
0
, then we have
|x x
0
|δ = |f (x) f (x
0
)|
δ
p
x
0
ε
=
p
x is continuous at x
0
.
Proposition 3.6.3. Let f , g : (a,b) 7→R continuous at x
0
(a,b). Then
(i) f +g is continuous at x
0
(ii) For λ R, λ · f is continuous at x
0
(iii) f ·g is continuous at x
0
(iv) f 6=0 on (a,b),
g
/f is continuous at x
0
(v) f : (a, b) 7→(c,d), g : (c, d) 7→R, f is continuous at x
0
(a,b) and g is
continuous at f (x
0
) (c,d), then g f is continuous at x
0
.
Proof. Left as an exercise.
30
Example. Let f : (0,+∞) 7→ R. We claim f (x) =
p
x is continuous at every
x
0
(0,+∞).
Consider f : R 7→R defined by
f (x) =
(
1 x Q
0 x R \ Q
which is discontinuous everywhere. Then let x
0
R,n N. We know that there
exists a
n
,b
n
(x
0
1
/n, x
0
+
1
/n) such that a
n
Q,b
n
R\Q such that a
n
,b
n
n→∞
x
0
. Then we have
f (a
n
) =1 1
f (b
n
) =0 0
= f discontinuous at x
0
.
Definition 3.6.4 (Continuity on an interval). We say f : (a,b) 7→R is continu-
ous on (a,b) if it is continuous at every x
0
(a,b).
We write f C
0
¡
(a,b)
¢
.
Definition 3.6.5 (Left/right limits). Let f : (a,b)\{x
0
} 7→R. We say that f has a
left-sided limit l R at x
0
if ε >0 δ(x
0
,ε) >0 such that
0 <|x x
0
|δ
|
{z }
xδxx
0
= |f (x) l |ε.
We write lim
xx
0
f (x) =x
0
.
Similarly, we say that f has a right-sided limit l R at x
0
if ε >0 δ(x
0
,ε) >0
such that
0 <|x x
0
|δ
| {z }
x
0
xx
0
+δ
= |f (x) l |ε.
We write lim
xx
+
0
f (x) =x
0
.
Definition 3.6.6 (Left/right continuous). We say that f : (a,b) 7→ R is left-
continuous at x
0
(a,b] if lim
xx
0
f (x) = f (x
0
) ε >0 δ >0 such that
x
0
δ x x
0
= |f (x) l |ε.
31
Similarly, we say that f : (a,b) 7→R is right-continuous at x
0
[a,b) if lim
xx
+
0
f (x) =
f (x
0
) ε >0 δ >0 such that
x
0
x x
0
+δ = |f (x) l|ε.
Example. Let f (x) =
p
x. Is f (x) is right-continuous at 0? i.e., lim
xx
+
0
f (x)
?
=0.
We note that f (x) is continuous on (0,1) but does not have a right limit at 0.
Lemma 3.6.7. Let f : (a,b) 7→ R, x
0
(a,b). Then f is continuous at x
0
if
and only if
lim
xx
0
f (x) = lim
xx
+
0
f (x) = f (x
0
).
Proof. Left as an exercise.
Definition 3.6.8 (Continuity on a closed interval). We say f : [a, b] 7→R is con-
tinuous on a closed interval if it is continuous on (a,b) and lim
xx
+
0
f (x) =
f (a),lim
xx
0
f (x) = f (b).
We write f C
0
¡
[a,b]
¢
.
Theorem 3.6.9 (Intermediate Value Theorem). Let f : (a,b) 7→R be a con-
tinuous function such that f (a) < f (b). Then for every y
0
R such that
f (a) < y
0
< f (b), there exists x
0
(a,b) such that f (x
0
) = y
0
(symmetrical
for f (b) < f (a)).
BCaution: y
0
may not be unique.
Proof. Let X ={x [a, b] | f (x) < y
0
} [a,b]. We know that X 6=; since a X
by definition. Further, since X is a subset of a bounded set, we know that X is
bounded. Therefore, we know that s =sup(X ) exists.
Now, we want to show
(i) a < s <b and
(ii) f (x) = y
0
.
If we can show these two things then we get that x
0
=s.
(i) We want to show that a < s. Recall that f is right-continuous at a means
that ε >0 δ >0 such that
a x a +δ = |f (x) f (a)|ε.
32
In particular, for e =(y
0
f (a))/2, we obtain
a x a +δ = |f (x) f (a)|ε =
y
0
f (a)
2
= f (x) f (a) +ε =
f (a)+y
0
2
< y
0
by construction. Now for the sake of contradiction, assume that sup(X ) =
s = a. Then take z =
(a+δ)
/2 and f (z) < y
0
by construction which fur-
ther implies that z X . However, z > a = s which contradicts the initial
assumption that s = sup(X ). Therefore, our contradiction assumption
must be false and a <s.
Next, we must show that s <b. Recall that since f is left-continuous at b,
we have ε >0 δ >0 such that
b δ x b = |f (x) f (b)|ε = f (x) f (b) ε.
Therefore, take ε = f (b) y
0
= f (x) y
0
by construction. For the
sake of contradiction, assume that sup(X ) = s = b. Then we have that
bδ x b = x 6∈ X . So take z =b
δ
/2 which means that z x x X .
However, this contradicts the initial assumption that s is the smallest up-
per bound of X . Therefore our contradiction assumption must be false
and s <b.
(ii) We want to show that f (s) = y
0
. First, we will show that f (s) y
0
. Let
(z
n
)
nN
, z
n
(s,s +
1
/n), z
n
s. Then we surely have that z
n
> s,z
n
6∈ X
which implies that (by the definition of X ) f (z
n
) y
0
. Note that by com-
parison theorems, we have that lim
n→∞
f (z
n
) y
0
(since f (z
n
) y
0
n
N). Also, since f is continuous on the interval, we have
y
0
lim
n→∞
f (z
n
) = f
¡
lim
n→∞
z(n)
¢
= f (s) = f (s) y
0
.
Next, we will show that f (s) y
0
. Recall the property of supremum:
z
n
X such that z
n
s. Then we know that f (z
n
) < y
0
since z
n
X .
Similarly to before, we have
y
0
lim
n→∞
f (z
n
) = f
³
lim
n→∞
f (z
n
)
´
= f (s) = f (s) y
0
.
Together, (i) and (ii) imply that f (s) = y
0
.
BWarning: If f : [a,b] 7→R such that f (a) < f (b) and y
0
R with f (a) < y
0
<
f (b) = x
0
(a,b) with f (x
0
) = y
0
6= f is continuous on [a,b]!
33
Definition 3.6.10 (Monotone functions). A function f : [a,b] 7→ R is mono-
tone increasing if
x, y [a,b] with x y = f (x) f (y).
Similarly, such a function is monotone decreasing if
x, y [a,b] with x y = f (x) f (y).
Lemma 3.6.11. Let f : [a,b] 7→ R be monotone and the conditions of the
Intermediate Value Theorem hold. Then f is continuous.
Proof. Details of proof omitted.
Corollary. Let f : [a, b] 7→R be a continuous function such that f (a)· f (b) 0.
Then there exists x
0
[a,b] such that f (x
0
) =0.
Proof. There are two cases to consider:
(1) Either f (a) =0 or f (b) =0.
(2) Both f (a) 6=0 and f (b) 6=0 = f (a) · f (b) <0.
Without loss of generality, assume f (a) < 0 and f (b) > 0. Then by the Inter-
mediate Value Theorem, y
0
=0 and f (a) < y
0
< f (b) and x
0
(a,b) such that
f (x
0
) = y
0
=0.
Corollary (Fixed Point). Let f : [a,b] 7→ [a,b] be a continuous function. Then
there exists x
0
[a,b] such that x
0
= f (x
0
).
Proof. Let g (x) =x f (x) be continuous on [a, b]. Then we have
g (a) =a f (a) 0
g (b) =b f (b) 0
since f : [a,b] 7→[a,b]. Then, by the previous corollary we have
g (a) ·g (b) 0 = x
0
[a,b] g (x) =0 = x
0
= f (x
0
).
Definition 3.6.12 (Boundedness on a set). Let X R be non-empty. A func-
tion f : X 7→R is said to be bounded on X iof there is M 0 such that
|f (x)| M x X .
34
Lemma 3.6.13. Let X R be non-empty. If f : X 7→R is not bounded on X ,
then (z
n
)
nN
X such that |f (z
n
)|n.
Proof. Boundedness implies that M 0 such that |f (x)| M x X so by
negation unboundedness implies M 0 x X such that |f (x)| > M. Then
consider m =n,(z
n
)
nN
X and we have |f (z
n
)|n.
Theorem 3.6.14 (Extreme Value Theorem). Let f : [a,b] 7→ R be a contin-
uous function. Then f is bounded on [a,b] and there are x
i
, x
g
[a,b] such
that
sup{f (x) : x [a,b]} = f (x
g
)
inf{f (x) : x [a,b]} = f (x
i
).
We denote these as
sup
x[a,b]
f (x) =sup{f (x) : x [a,b]}
inf
x[a,b]
f (x) =inf{f (x) : x [a,b]}.
Proof. We first must show that f is bounded on [a,b].
For the sake of contradiction, assume that f is unbounded (not bounded).
Then there exists a sequence (z
n
)
nN
[a,b] with |f (z
n
)|n = f divergent.
However, (z
n
)
nN
is bounded and Bolzano-Weirstrass implies that (z
n
k
)
kN
such that z
n
k
l [a,b]. Further, by comparison results we have
lim
k→∞
f (z
n
k
) = f ( lim
k→∞
z
n
k
) = f (l ).
But we know that ( f (z
n
k
))
kN
converges to f (l ) which implies that f is con-
vergent which is a contradiction. Therefore f must be unbounded.
Now we must show that f bounded = X bounded = M =sup(X ) and m =
inf(X ) exists. So we need to show that M = f (x
g
), x
g
[a,b] and m = f (x
i
), x
i
[a,b]. From the characterization of infimum, we know ε > 0 q
²
X such
that q
ε
m ε. Then we have that q
ε
X f (x
ε
), x
ε
[a,b]. This means that
f (x
ε
) m ε. Taking ε =
1
/n we obtain m f (x
n
) m +
1
/n = f (x
n
) m by
the Squeeze Theorem.
But (x
n
)
nN
bounded implies there exists some subsequence (x
n
k
)
kN
such
35
that x
n
k
[a,b] = x
n
k
l [a,b]. So
m = lim
k→∞
f (x
n
k
) = f ( lim
k→∞
x
n
k
) = f (l ).
Take x
i
=l [a,b].
Definition 3.6.15 (Supremum and max of a function). Let f : X R, X non-
empty. If there is x X such that f (x) =sup{f (x) : x X }, we write sup
xX
f (x) =
max
xX
f (x). Similarly for inf.
For f : [a,b] 7→R, the EVT guarantees that sup
x[a,b]
f (x) =max
x[a,b]
f (x) and
inf
x[a,b]
f (x) =min
x[a,b]
f (x).
3.7 Uniform Continuity
Remark. If (z
n
)
nN
is bounded and f is a continuous function, then (f (z
n
))
nN
is bounded.
Remark. However, (z
n
)
nN
convergent 6= ( f (z
n
))
nN
convergent!
Example. Consider f (x) =
1
/x which is continuous on (0,1). Indeed, x
0
(0,1), ε >0 δ =δ(ε, x
0
) such that
|x x
0
|δ = |f (x) f (x
0
)|ε
¯
¯
¯
¯
1
x
1
x
0
¯
¯
¯
¯
=
¯
¯
¯
¯
x x
0
x ·x
0
¯
¯
¯
¯
4|x x
0
|
3x
2
0
ε ( by (*))
since
x
0
δ x x
0
+δ
δ
x
0
4
= x
3
4
x
0
(*)
However, since the definition of continuity is dependent on x
0
which is not
constant, we have a problem. Take z
n
=
1
/n 0. But f is continuous on (0,1).
So f (z
n
) =
1
/(1/n) = n is not convergent.
Definition 3.7.1 (Uniform continuity). Let X R be non-empty. We say that
f : X 7→R is uniformly continuous on X if ε >0 δ =δ(ε) such that |X y|
δ = |f (x) f (y)|ε.
36
Lemma 3.7.2. Let f : X 7→ R, X R non-empty where f is uniformly con-
tinuous. Then f is continuous.
Proof. To show f is continuous at x
0
X , we need to show that ε > 0 δ =
δ(ε, x
0
) >0 such that
|x x
0
|δ = |f (x) f (x
0
)|ε.
Since f is assumed to be uniformly continuous, we know that
e
ε >0
e
δ =
e
δ(
e
ε)
such that
|x y|
e
ε = |f (x) f (y)|
e
ε.
Then let ε >0 be given. Let
e
ε =ε and
e
δ =
e
δ(
e
ε) and write uniformly continuous
for y =x
0
. Then
|x x
0
|
e
δ = |f (x) f (x
0
)|
e
ε =ε.
It suffices to take δ =
e
δ.
Example (re-visited). Consider f (x) =
1
/x which is uniformly continuous on
(1,2). Because ε >0, x, y (1, 2),
|f (x) f (y)|=
¯
¯
¯
¯
1
x
1
y
¯
¯
¯
¯
=
|x y|
x y
|x y|.
Let δ =ε, then clearly
|x y|δ = |f (x) f (y)|ε.
Now, let us show that f is not uniformly continuous on (0,1).
Remark. A function f : X 7→ R, X R non-empty is not uniformly continu-
ous on X if ε
0
> 0 and (x
n
)
nN
,(y
n
)
nN
X with |x
n
y
n
|
1
/n and |f (x
n
)
f (y
n
)|ε
0
.
So, coming back to the example, let x
n
=
1
/n, y
n
=
1
/(n+1). Then compute
|x
n
y
n
|=
1
n(n +1)
1
n
|f (x
n
) f (y
n
)|=|n (n +1)|=1.
In other words, the difference between any terms of the same index of the
image sequences is 1 while the difference between any two terms of the same
index in the domain sequence tends to 0.
37
Proposition 3.7.3. Let f : (a,b) 7→ R be uniformly continuous. If (z
n
)
nN
is Cauchy with z
n
(a,b), then (f (z
n
))
nN
is Cauchy.
Proof. Since f is uniformly continuous, we have that
e
ε >0
e
δ =
e
δ(
e
ε) >0 such
that
|x y|
e
δ = |f (x) f (y)|
e
ε x, y (a,b).
Also, since (z
n
)
nN
Cauchy, we have that ε >0 N N such that m,n N ,
|z
m
z
n
|ε.
Next, we need to show that ε > 0 N N such that m,n N , |f (z
n
)
f (z
m
)| ε. Let ε > 0 be given. Set
e
ε = ε which implies that there exists
e
δ > 0
such that
|x y|
e
δ = |f (x) f (y)|ε.
Then, set ε =
e
δ = N N such that m,n N,|z
n
z
m
| ε =
e
δ. Finally, set
N = N and we get that, for all m,n N,
|f (z
n
) f (z
m
)|ε.
Theorem 3.7.4 (Uniform continuity). Let f : [a,b] 7→ R be a continuous
function. Then f : [a,b] 7→R is uniformly continuous.
Proof. For the sake of contradiction, assume f is not uniformly continuous.
Then ε
0
>0 and sequences (x
n
)
nN
,(y
n
)
nN
such that x
n
, y
n
[a,b] with
|x
n
y
n
|
1
n
and |f (x
n
) f (y
n
)|ε
0
.
Further, since x
n
[a,b], we know that x
n
is bounded and by the Bolzano-
Weirstrass Theorem, there exists a subsequence (x
n
k
)
kN
such that x
n
k
l
1
[a,b].
Next, consider another subsequence (y
n
k
)
kN
which not necessarily conver-
gent (this is because, we are taking the indices the same as x
n
k
so we know the
subsequence exists but not that it converges). However, since y
n
k
is bounded,
the Bolzano-Weirstrass Theorem implies that it also has a convergent sub-
subsequence, say y
m
k
of the subsequence y
n
k
such that y
m
k
l
2
[a,b]. Also,
we can similarly say that (x
m
k
)
kN
is a subsequence of x
n
k
so x
m
k
l
1
.
We now must show that l
1
= l
2
. Consider |x
m
k
y
m
k
|
1
/m
k
. As k , the
38
difference (x
m
k
y
m
k
) 0 and
1
/m
k
0. Therefore, lim
k→∞
x
n
k
= l
1
= l
2
=
lim
k→∞
y
m
k
.
Now, we know f is continuous, so lim
k→∞
f (x
m
k
) = f (l
1
) and lim
k→∞
f (y
m
k
) =
f (l
2
) and l
1
= l
2
. Together, these two things imply that f (l
1
) = f (l
2
). This
means that lim
k→∞
|f (x
m
k
y
m
k
)|= 0 which contradicts the assumption that
|f (x
m
k
) f (y
m
k
)| ε
0
. Therefore, our initial assumption is false and f must
be uniformly continuous.
Theorem 3.7.5 (Continuity Extension). Let f : (a,b) 7→ R be a function.
Then f is uniformly continuous on (a,b) if and only if there is some con-
tinuous function g : [a,b] 7→R such that g (x) = f (x) on (a,b).
Proof. First, assume that g : [a,b] 7→ R is continuous. From Theorem 3.7.4,
g is uniformly continuous which implies that g is uniformly continuous on
(a,b) [a,b] which implies that f is continuous on (a,b) since f = g .
Next, assume that f is uniformly continuous on (a,b). Take g(x) = f (x) when-
ever x (a,b), i.e., g is also continuous on (a,b). Now, for g to be continuous
on [a,b], we need to decide the values of g (a) and g (b). In particular, it must
be that
g (a) = lim
xa
+
g (x) = f (x) and g (b) = lim
xb
g (x) = f (x)
by the definition of right and left continuity. By symmetry, we will only show
that g (a) =lim
xa
+
g (x) = f (x), and showing the g (b) case is nearly identical.
By the characterization of sequential limits, we have that lim
xa
+
g (x) =l ex-
ists and is finite if for every sequence (z
n
)
nN
, z
n
(a,b), we have
z
n
a and f (z
n
) l.
Take (z
n
)
nN
, z
n
(a,b),z
n
a. Since (z
n
)
nN
is converging to a limit, it is
Cauchy. Now since f is uniformly continuous and z
n
is Cauchy, Proposition
3.7.3 tells us that (f (z
n
))
nN
is also Cauchy and is thus convergent to some
limit, say l
z
. We now show that the limit l
z
is independent of the sequence.
For the sake of contradiction, suppose there is some sequence (y
n
)
nN
, y
n
(a,b), y
n
a, then we know that f (y
n
) l
y
6=l
z
. Further, our initial assump-
tions yield some properties which we will list below:
(1) f (z
n
) l
z
means that ε
1
>0 N
1
N such that n N
1
|f (z
n
) l
z
|ε
1
.
(2) f (y
n
) l
y
means that ε
2
>0 N
2
N such that n N
2
|f (y
n
) l
z
|ε
2
.
39
(3) f being uniformly continuous means that ε
3
> 0 δ
3
> 0 such that, for
x, y (a,b)
|x y|δ
3
= |f (x) f (y)|ε
3
.
(4) z
n
a means that ε
4
>0 N
4
N such that n N
4
|z
n
a|ε
4
.
(5) y
n
a means that ε
5
>0 N
5
N such that n N
5
|y
n
a|ε
5
.
Let ε > 0 be given. Further, set ε
1
=
ε
/3,ε
2
=
ε
/3,ε
3
=
ε
/3 = δ
3
by (3). Then,
set ε
4
=
δ
3
/2,ε
5
=
δ
3
/2. Why make these particular choices? Because then we
have
|z
n
a|
δ
3
2
n N
4
, (by (4))
|y
n
a|
δ
3
2
n N
5
, (by (5))
=|z
n
a y
n
+a||z
n
a|+|y
n
a|
=
δ
3
2
+
δ
3
2
=δ
3
= |z
n
y
n
|δ
3
n max{N
4
, N
5
}.
Then, by (3), this implies that |f (z
n
) f (y
n
)|
ε
/3. Therefore, by the triangle
inequality we get
|l
z
l
y
||l
z
f (z
n
)|
| {z }
ε
/3 nN
1
+ |f (z
n
) f (y
n
)|
| {z }
ε
/3 nmax{N
4
,N
5
}
+|f (y
n
) l
y
|
| {z }
ε
/3 nN
2
ε
3
+
ε
3
+
ε
3
=ε.
So |l
z
l
y
| ε,n max{N
1
, N
2
, N
4
, N
5
} ε > 0 = l
z
= l
y
. However, this con-
tradicts our assumption that l
z
6= l
y
, therefore l
z
= l
y
and we can set g (a) =
l
z
=l
y
.
3.8 Inverses and Continuity
Example. Take the function f (x) =sin(x), which is continuous on the interval
(
π
/2,
π
/2).
40
π/2 π/2
1
1
sin(x)
x
f (x)
1 1
π/2
π/2
sin
1
(x)
f (x)
x
Although the picture makes it look obvious, we will rigorously show that the
inverse of a continuous function is also continuous.
Proposition 3.8.1. Let I R be an interval (could be closed, open, etc.). If
f : I 7→R is injective and continuous, then f is strictly increasing or strictly
decreasing.
41
Proof. We will show for the case where I = (a,b) and the other cases are the
same. For the sake of contradiction, assume that there are
a < x
1
<x
2
<x
3
<b
such that
f (x
1
) f (x
2
) and f (x
3
) f (x
2
),
i.e., not strictly increasing. Since f is injective, f (x
1
) f (x
2
) = f (x
1
) < f (x
2
)
and f (x
3
) f (x
2
) = f (x
3
) < f (x
2
). Therefore, α R such that
f (x
1
) <α < f (x
2
) and f (x
3
) <α < f (x
2
).
Now, since f is continuous on [x
1
, x
2
] and [x
2
, x
3
], the Intermediate Value The-
orem (3.6.9) implies that
e
x (x
1
, x
2
) such that f (
e
x) =α and
e
e
x (x
2
, x
3
) such
that f (
e
e
x) =α. By the conditions of the IVT, it is clear that
e
x 6=
e
e
x. But this gives
us that f (
e
x) = f (
e
e
x) which contradicts the fact that f is injective. Therefore,
our contradiction assumption must be false and f is either strictly increasing
or strictly decreasing.
Theorem 3.8.2 (Inverse continuity). Let f : [a,b] 7→ R be continuous and
injective. Then f ([a,b]) is a closed interval I and f
1
: I 7→[a, b].
Proof. Never got around to this in class.
4 Differentiation
Definition 4.0.1 (Locally differentiable). Let f : (a,b) 7→ R and x
0
(a,b). If
lim
xx
0
(f (x) f (x
0
))/(x x
0
) exists and is finite, then we say that f is locally
differentiable at x
0
and write
f
0
(x
0
) = lim
xx
0
f (x) f (x
0
)
x x
0
.
Remark.
(1) By function composition and taking g (x) = x
0
+h, we note that
lim
xx
0
f (x) f (x
0
)
x x
0
= lim
h0
f (x
0
+h) f (x
0
)
h
.
(2) Since f is differentiable at x
0
, it is also continuous at x
0
. This is because
you can always write
f (x) = f (x
0
) +
f (x) f (x
0
)
x x
0
(x x
0
) (x 6= x
0
)
42
Definition 4.0.2 (Globally differentiable). Let f : (a,b) 7→ R. We say f is glob-
ally differentiable on the interval (a,b) if f is differentiable on every x
0
(a,b). In that case, we have
x 7→ lim
xx
f (x) f (x)
x x
= f
0
(x)
is a well-defined function on (a,b) called the derivative of f on (a,b).
Examples.
(1) Let f (x) =α. We claim this is differentiable on R.
Let x
0
R, x
0
6=x. Compute
f (x) f (x
0
)
x x
0
=
α α
x x
0
=0.
So lim
xx
0
f (x)f (x
0
)
xx
0
exists and is finite and equals 0.
(2) Let f (x) = x
3
. We claim this is differentiable on R. Let x
0
R, x
0
6= x.
Compute
f (x) f (x
0
)
x x
0
=
x
3
x
3
0
x x
0
=x
2
+x
0
x +x
2
0
.
So lim
xx
0
f (x)f (x
0
)
xx
0
exists and is finite and equals 3x
2
0
.
(3) Let f (x) = |x|. We claim this is differentiable on (−∞,0) (0,) but not
at 0. First, consider x
0
>0. Subsequently, f (x) =x
0
>0. Then compute
f (x) f (x
0
)
x x
0
=
x x
0
x x
0
=1 = f
0
(x
0
) =1.
Similarly, for x
0
<0, f
0
(x
0
) =1. When x
0
=0, we have
lim
xx
0
f (x) f (x
0
)
x x
0
= lim
xx
0
|x|
x
.
As previously established (similar to Example 3.1), this limit does not ex-
ist, so f is not differentiable at 0.
(4) Let f (x) =sin(x). We claim this is differentiable on R. Take x
0
R. Com-
pute
f (x) f (x
0
)
x x
0
=
sin(x) sin(x
0
)
x x
0
=
1
2
sin(
xx
0
2
) ·cos(
x+x
0
2
)
x x
0
=
sin(
xx
0
2
)
xx
0
2
·cos(
x +x
0
2
).
Then, we have
lim
xx
0
Ã
sin(
xx
0
2
)
xx
0
2
·cos
³
x +x
0
2
´
!
=
lim
xx
0
sin(
xx
0
2
)
lim
xx
0
xx
0
2
· lim
xx
0
cos(
x +x
0
2
)
=1 ·cos(x
0
).
43
Since this limit exists and is finite, we have f
0
(x
0
) = cos(x
0
). Similarly,
f (x) =cos(x) = f
0
(x
0
) =sin(x
0
).
Theorem 4.0.3 (Caratheodory’s Theorem). Let f : (a,b) 7→ R, x
0
(a,b).
The function f is differentiable at x
0
if and only if there is ϕ : (a,b) 7→R that
is continuous at x
0
and f (x) = f (x
0
)+ϕ(x)(xx
0
) x (a,b). Furthermore,
f
0
(x
0
) =ϕ(x
0
).
Proof. First, suppose such a ϕ exists. Further, suppose x 6= x
0
. Then we have
f (x) f (x
0
)
x x
0
=ϕ(x).
Furthermore,
lim
xx
0
f (x) f (x
0
)
x x
0
= lim
xx
0
ϕ(x) =ϕ(x
0
) ϕ continuous at x
0
f
0
(x
0
) =ϕ(x
0
).
Next, suppose f
0
(x
0
) is well-defined. In particular, set
ϕ =
(
f (x)f (x
0
)
xx
0
x 6= x
0
f
0
(x
0
) x = x
0
.
Clearly, ϕ(x) is continuous at x
0
, so
lim
xx
0
ϕ(x) = lim
xx
0
f (x) f (x
0
)
x x
0
= f
0
(x
0
) =ϕ(x
0
).
On the other hand, when x 6= x
0
, we have
ϕ(x) =
f (x) f (x
0
)
x x
0
f (x) = f (x
0
) +ϕ(x
0
)(x x
0
)
f (x) = f (x
0
) +ϕ(x)(x x
0
).
Lemma 4.0.4. Let f : (a,b) 7→ R, If f is differentiable at x
0
(a,b), then f
is continuous at x
0
.
Proof. If f is differentiable at x
0
, then there exists ϕ which is continuous at
x
0
such that f (x) = f (x
0
) +ϕ(x)(x x
0
) and lim
xx
0
f (x) = f (x
0
) +ϕ(x
0
) ·0 =
f (x
0
).
44
Remark. There exists functions whose derivatives are not continuous. For
example, consider
f (x) =
(
x
2
sin
¡
1
x
¢
x 6= 0
0 x = 0.
4.1 Extremum
Definition 4.1.1 (Local extremum). Let f : I 7→ R. We say that c I is a local
minimum/maximum if there is δ >0 such that f (x) / c x (c δ,c +δ).
An extremum is a minimum or a maximum.
Theorem 4.1.2 (Fermat). Let f : (a,b) 7→ R. If f has an extremum at c
(a,b) and is differentiable at c, then f
0
(c) =0.
Proof (Minimum). Let c be such an element. We have c is a local minimum
which implies that there exists δ >0 such that f (x) c x (c δ,c +δ). Take
x (c, c +δ) and compute the ration
f (x) f (c)
x c
0 (x bigger than c by construction)
comp.
lim
xc
+
f (x) f (x)
x c
0.
Similarly, for x (c δ,c), we have
lim
xc
f (x) f (c)
x c
0.
Next, we have that f is differentiable at c which implies
0 lim
xc
+
f (x) f (c)
x c
= lim
xc
f (x) f (c)
x c
= lim
xc
f (x) f (c)
x c
0
lim
xc
f (x) f (c)
x c
=0 = f
0
(c).
4.2 Rules of Differentiation
4.2.1 Basic Rules
45
Proposition 4.2.1 (Basic rules of differentiation). Let f ,g : (a,b) 7→ R dif-
ferentiable at x
0
(a,b), λ R.
(i) f +g is differentiable at x
0
with
(f +g )(x
0
) = f
0
(x
0
) +g
0
(x
0
).
(ii) λ · f is differentiable at x
0
with
(λ · f )
0
=λ · f
0
(x
0
).
Proof. (i) Take x
0
(a,b) and compute
lim
xx
0
(f +g )(x) (f +g)(x
0
)
x x
0
= lim
xx
0
f (x) +g (x) f (x
0
) g (x
0
)
x x
0
= lim
xx
0
f (x) f (x
0
)
x x
0
+ lim
xx
0
g (x) g(x
0
)
x x
0
= f
0
(x
0
) +g
0
(x
0
).
(ii) Compute
lim
xx
0
(λf )(x)(λf )(x
0
)
x x
0
= lim
xx
0
λ
f (x) f (x
0
)
x x
0
=λ · lim
xx
0
f (x) f (x
0
)
x x
0
=λ · f
0
(x
0
).
4.2.2 Advanced Rules
Proposition 4.2.2 (Product Rule). Let f ,g : (a,b) 7→R be differentiable at
x
0
(a,b). Then f ·g is differentiable at x
0
with
(f ·g )
0
(x
0
) = f
0
(x
0
)g (x
0
) + f (x
0
)g
0
(x
0
).
46
Proof. Let x (a,b)\{x
0
} and compute
lim
xx
0
(f ·g )(x) (f ·g)(x
0
)
x x
0
= lim
xx
0
f (x)g (x) f (x
0
)g (x
0
)
x x
0
= lim
xx
0
¡
f (x) f (x
0
)
¢
g (x) + f (x
0
)
¡
g (x) g(x
0
)
¢
x x
0
= lim
xx
0
·
f (x) f (x
0
)
x x
0
·g (x) +
g (x) g(x
0
)
x x
0
· f (x
0
)
¸
= lim
xx
0
f (x) f (x
0
)
x x
0
·g (x) + lim
xx
0
g (x) g(x
0
)
x x
0
· f (x
0
)
= f
0
(x
0
) ·g (x
0
) + f (x
0
) ·g
0
(x
0
). (*)
(*) Note: Since g , f are differentiable at x
0
, they are also continuous at x
0
and
thus lim
xx
0
g (x), f (x) =g (x
0
), f (x
0
).
Proposition 4.2.3 (Quotient Rule). Let f : (a,b) 7→ R be differentiable at
x
0
(a,b) such that f (x
0
) 6=0. Then there exists δ >0 such that f (x) 6=0 for
all x (x
0
δ, x
0
+δ) (f well-defined on (x
0
δ, x
0
+δ)) and
µ
1
f
0
(x
0
) =
f
0
(x
0
)
f (x
0
)
2
.
Then, assume g (x
0
) 6=0. Then
f
/g is differentiable at x
0
and
µ
f
g
0
(x
0
) =
f
0
(x
0
)g (x
0
) f (x
0
)g
0
(x
0
)
(g (x
0
))
2
.
Proof. We will first show the first part of the proposition and the second part
47
follows immediately. Take x (x
0
δ, x
0
+δ) and compute
lim
xx
0
1
f
(x)
1
f
(x
0
)
x x
0
= lim
xx
0
1
f (x)
1
f (x
0
)
x x
0
= lim
xx
0
f (x)f (x
0
)
f (x)·f (x
0
)
x x
0
= lim
xx
0
·
f (x) f (x
0
)
x x
0
·
1
f (x) · f (x
0
)
¸
= lim
xx
0
f (x) f (x
0
)
x x
0
· lim
xx
0
1
f (x) · f (x
0
)
=f
0
(x
0
) ·
1
(f (x
0
))
2
=
f
0
(x
0
)
(f (x
0
)
2
)
.
For the second part, we note that
f
/g = f ·
1
/g and compute
µ
f
g
0
(x
0
) = f
0
(x
0
) ·
1
g (x
0
)
+ f (x
0
) ·
µ
g
0
(x
0
)
(g (x
0
))
2
| {z }
product rule + quotient rule
=
f
0
(x
0
)g (x
0
) + f (x
0
)g
0
(x
0
)
(g (x
0
))
2
.
Proposition 4.2.4 (Chain Rule). Let f : (a,b) 7→ (c,d) be differentiable at
x
0
(a,b) and let g : (c, d) 7→R be differentiable at f (x
0
) (c,d). Then g f
is differentiable at x
0
with
(g f )
0
(x
0
) =g
0
(f (x
0
))f
0
(x
0
).
Proof. Recall Caratheodory’s Theorem (4.0.3) which states that a function f is
differentiable at x
0
if and only if there exists a continuous function ϕ : (a,b) 7→
R such that f (x) = f (x
0
) +ϕ(x)(x x
0
) and ϕ(x
0
) = f
0
(x
0
). We can tweak this
equation for g as follows:
The function g is differentiable at f (x
0
) if and only if there exists a
continuous function Φ : (c,d) 7→R such that g (y) =g (f (x
0
)) +Φ(y)(y f (x
0
))
and Φ(f (x
0
)) =g
0
(f (x
0
)).
Now compute for x (a,b)\{x
0
}.
(g f )(x) (g f )(x
0
)
x x
0
=
g (f (x)) g( f (x
0
))
x x
0
=
Φ(f (x))(f (x) f (x
0
))
x x
0
.
48
Now,
lim
xx
0
Φ(f (x)) · lim
xx
0
f (x) f (x
0
)
x x
0
=Φ(f (x
0
)) · f
0
(x
0
)
=g
0
(f (x
0
)) · f
0
(x
0
)
since Φ is continuous and f is differentiable at x
0
. All of this together implies
that (g f )(x
0
) =g
0
(f (x
0
)) · f
0
(x
0
).
4.3 Mean Value Theorem and Applications
Theorem 4.3.1 (Rolles Theorem). Let f : [a, b] 7→R be continuous on [a, b]
and differentiable on (a,b). If f (a) = f (b), then there exists c (a,b) such
that f
0
(c) =0.
Proof. First note that f continuous on [a,b] implies (by the Extreme Value
Theorem) that there exists r, s [a, b] such that f (r )
|{z}
m
f (x) f (s)
|{z}
M
x [a,b].
Then consider three cases:
(1) If m = M, then f (x) = m = M x [a,b] which further implies that
f
0
(x) =0 x (a,b) taking any c (a,b).
(2) If m < M, then either r (a,b) or s (a,b) because f (a) = f (b) by as-
sumption.
(i) If r (a,b), then f achieves its minimum at r and is thus differen-
tiable at r , then Fermats Theorem (4.1.2) implies that f
0
(r ) =0, tak-
ing c =r we get our desired result.
(ii) Take s (a,b) and the same result is obtained.
(3) The case where m >M is symmetric and left as an exercise.
Theorem 4.3.2 (Mean Value Theorem). Let f : [a,b] 7→ R be continuous
on [a,b] and differentiable on (a,b). Then there exists a c (a,b) such that
f
0
(c) =
f (b) f (a)
b a
.
Proof. Define h(x) = f (x)(ba)+(a x)(f (b)f (a)). Since h is a composition
of two differentiable and continuous functions, h is continuous on [a,b] and
49
differentiable on (a, b). Then compute
h(a) = f (a) ·b f (a) ·a
h(b) = f (b)(b a) (b a)(f (b) f (a))
=(b a) · f (a) =h(a)
Then Rolles Theorem (4.3.1) implies that there exists a c (a,b) such that
h
0
(c) =0 which means that
h
0
(x) = f
0
(x)(b a) (f (b) f (a))
0 =h
0
(c) = f
0
(c)(b a) (f (b) f (a))
f
0
(c) =
f (b) f (a)
b a
.
Theorem 4.3.3 (Mean Value Theorem (Cauchy)). Let f ,g : [a, b] 7→ R be
continuous on [a,b] and differentiable on (a,b). Further assume g
0
(x) 6=
0 x (a,b). Then there exists c (a,b) such that
f
0
(c)
g
0
(c)
=
f (b) f (a)
g (b) g(a)
.
Proof. First, note that g(a) 6=g (b) because otherwise, Rolles Theorem implies
that there exists some d (a, b) such that g
0
(d) = 0, which is a contradiction.
Thus, our ratio is well-defined.
Let h(x) = f (x)(g (b)g (a))(g (x)g(a))(g (b) f (a)). As before, h us contin-
uous and differentiable on [a,b]/(a,b), respectively. Then
h(a) = f (a)g (b) f (a)g (a)
h(b) = f (b)(g (b) g(a))(g (b) g(a))(f (b) f (a))
=g (b)f (a) g (a)f (a) =h(a).
Then Rolles Theorem implies that there exists some c (a,b) such that h
0
(c) =
0 and we have
h
0
(x) = f
0
(x)(g (b) g(a))(g
0
(x)(f (b) f (a)))
0 =h
0
(c) = f
0
(c)(g (b)g (a)) g
0
(c)(f (b) f (a))
f
0
(c)
g
0
(c)
=
f (b) f (a)
g (b) g(a)
.
50
Proposition 4.3.4 (L’Hospital’s Rule). Let f , g : (a, b) 7→R be two continu-
ous functions differentiable on (a,b) \ {x
0
} with
g
0
(x) 6=0 x (a,b)\{x
0
},
lim
xx
0
f (x) =lim
xx
0
g (x) =0,
and there is l R such that lim
xx
0
f
0
(x)/g
0
(x) =l.
Then
lim
xx
0
f (x)
g (x)
=l.
Proof. We show that lim
xx
+
0
f (x)/g (x) =l =lim
xx
0
f (x)/g (x). We first show
that g cannot vanish (become zero) more than once on the interval (a,x
0
). Let
a < x
1
<x
2
<x
0
.
a x
1
x
2
x
0
What can be said about g : [x
1
, x
2
] 7→ R? Certainly, g is continuous since it is
differentiable on the entire interval, so g is continuous on [x
1
, x
2
] and differ-
entiable on (x
1
, x
2
). Then, the MVT (Theorem 4.3.2) implies that there exists a
c (x
1
, x
2
) (a,b) such that
0 6=g
0
(c) =
g (x
2
) g (x
1
)
x
2
x
1
which implies that g cannot vanish at x
1
or x
2
.
Say that g does vanish on γ such that a <γ < x, if it doesnt vanish, then γ =a.
By assumption, we now that ε >0 δ > 0 such that if x (a,x
0
) and (WLOG)
γ <x
0
δ x x
0
, then
¯
¯
¯
¯
f
0
(x)
g
0
(x)
l
¯
¯
¯
¯
ε.
Let γ < x
0
δ x < y < x
0
. Then, since f and g are continuous on the entire
interval, the MVT Cauchy Version (Theorem 4.3.3) implies that there exists a
c (x, y) such that
f
0
(c)
g
0
(c)
=
f (y) f (x)
g (y)g (x)
.
γ
x
0
δ
x y x
0
c
51
Then consider
¯
¯
¯
¯
f
0
(x)
g
0
(x)
l
¯
¯
¯
¯
=
¯
¯
¯
¯
f (y) f (x)
g (y)g (x)
l
¯
¯
¯
¯
ε
comp.
= lim
yx
0
¯
¯
¯
¯
f
0
(x)
g
0
(x)
l
¯
¯
¯
¯
ε
=
¯
¯
¯
¯
f (x)
g (x)
l
¯
¯
¯
¯
ε
= lim
xx
0
f (x)
g (x)
=l.
Similarly, lim
xx
+
0
f (x)/g (x) =l.
4.4 Limited Expansion
Definition 4.4.1. For n N, we say that f : (a, b) 7→R is n-times differentiable
on (a,b) if
f is differentiable, f
0
is differentiable, .. ., f
(n)
is differentiable.
Theorem 4.4.2 (Taylor). Let f : (a,b) 7→R be (n +1) times differentiable (n
derivatives are continuous, but (n +1)
th
may not be). Then
f (x) =P
n
(x) +ε(x)(x a)
n
where
P
n
(x) = f (a) +
f
0
(a)
1!
(x a) +···+
f
(n)
(a)
n!
(x a)
n
and ε(x) =
f
(n+1)
(c)
/(n+1)!(x a) c (a,b).
Proof. Since P
n
is a polynomial of degree n in x, rewrite as follows:
P
n
(x) =
n
X
i=0
f
(i)
(a)
i!
(x a)
i
where f
(0)
= f .
Then
P
0
n
(x) =
n
X
i=1
f
(i)
(a)
i!
i ·(x a)
i1
=
n
X
i=1
f
(i)
(a)
(i 1)!
(x a)
i1
and similarly (for 1 j n)
P
(j )
n
(x) =
n
X
i=j
f
(i)
(a)
(i j )!
(x a)
ij
52
where P
(j )
n
(a) = f
(j )
(a) for j =0,...,n.
Let
g (x) = f (x) P
n
(x) +
P
n
(y) f (y)
(y a)
n+1
(x a)
n+1
y (a,b).
Then g is (n +1) times differentiable and
0 =g (a) =g
0
(a) =···=g
(n)
(a).
In addition, g (y) =0, so by Rolles Theorem (4.3.1), since
g (a) =g (y) =0, c
1
(a,b) such that g
0
(c
1
) =0
g
0
(a) =g
0
(c
1
) =0, c
2
(a,b) such that g
00
(c
2
) =0
.
.
.
g
(n)
(a) =g
(n)
(c
n
) =0, c (a,b) such that g
(n+1)
(c) =0.
Then we have
g
(n+1)
(x) = f
(n+1)
(x) +
P
n
(y) f (y)
(y a)
n+1
(n +1)!
0 = f
(n+1)
(c) +
P
n
(y) f (y)
(y a)
n+1
(n +1)!
f (y) =P
n
(y)+
f
(n+1)
(c)
(n +1)!
(y a)
n+1
| {z }
ε(y)(ya)
n
.
This is the result for x = y.
Example. Compute the Taylor Expansion for f (x) =sin(x).
sin(x) =0 +x
x
3
3!
+
x
5
5!
| {z }
P
5
(x)
+
sin(c)
6!
x
6
| {z }
ε(x)x
5
c (0, x).
Lemma 4.4.3 (Taylor Expansion uniqueness). Let f : (a,b) 7→R be (n +1)
times differentiable such that
f (x) =b
0
+b
1
(x a) +···+b
n
(x a)
n
+
e
ε(x)(x a)
n
b
i
R,
e
ε(x)
x0
0.
Then
b
i
=
f
(i)(a)
i!
,i =0,. .. ,n and
e
ε(x) =ε(x) =
f
(n+1)
(c)
(n +1)!
(y a) c (a,x).
53
Proof. We have two expansions of f :
(i) f (x) =
P
n
i=0
b
i
(x a)
i
+
e
ε(x)(x a) and
(ii) f (x) =
P
n
i=0
£
f
(i)
(a)/i!·(x a)
i
¤
+ε(x)(x a)
n
.
Consider f (x) f (x) =0
0 =
n
X
i=0
µ
f
(i)
(a)
i!
(x a)
i
+
(
ε(x)
e
ε(x)
)
(x a)
n
.
Assume by contradiction that
f
(k)
(a)
/k! 6= b
k
0 k n. Without loss of general-
ity, we take the smallest of such ks and get
0 =
n
X
i=k
µ
f
(i)
(a)
i!
b
i
(x a)
i
+(ε(x)
e
ε(x))(x a)
n
0 =
n
X
i=k
µ
f
(i)
(a)
i!
b
i
(x a)
ik
+(ε(x)
e
ε(x))(x a)
nk
(*)
= lim
xa
n
X
i=k
µ
f
(i)
(a)
i!
b
i
(x a)
ik
+(ε(x)
e
ε(x))(x a)
nk
=0 =
f
(k)
(a)
k!
b
k
.
Therefore,
f
(i)
(a)
/i! =b
i
i =0,...,n and by (*) we have
0 =(ε(x)
e
ε(x))(x a)
n
ε(x) =
e
ε(x).
Remark. Let f , g : (a, b) 7→R be (n +1) times differentiable functions such that
f (x) =
n
X
i=0
b
i
(x a)
i
+ε(x)(x a)
n
b
i
R, ε(x)
xa
0
g (x) =
n
X
i=0
d
i
(x a)
i
+ε(x)(x a)
n
d
i
R,
e
ε(x)
xa
0.
Then
(f +g ) =
n
X
i=0
(b
i
+d
i
)(x a)
i
+(ε(x) +
e
ε(x))(x a)
n
(f ·g ) =
n
X
i=0
(b
i
·d
i
)(x a)
i
+(ε(x) ·
e
ε(x))(x a)
n
(f /g ) =
n
X
i=0
(b
i
/d
i
)(x a)
i
+(ε(x)/
e
ε(x))(x a)
n
g 6=0.
54
Example. Consider the Taylor expansion of f (x) =
1
/(1+x). First realize that
(1 +x)(1 x +x
2
+···) =1.
The point is:
f (x) =
1
1 +x
=1 x +x
2
x
3
+···+) 1)
n
x
n
+ε(x)x
n
where ε(x) =(1)
n+1
x +(1)
n+2
x
2
+···. Then f
(n)
(0) =(1)
n
·n!.
Example. Consider
f (x) =
1
(1 +x
2
)
=1 x
2
+x
4
x
6
+···+(1)
n
x
2n
+ε(x)x
2n
.
Example. Consider f (x) =(sin
2
(x)x
2
cos(x))/(x
2
(1cos(x)) defined on (π/2,0)
(π/2,0). How can we compute lim
x0
f (x)? One could employ L’Hopital’s rule,
but this would be extremely messy. A slightly more convenient way to com-
pute this limit is the following. Let us consider the Taylor expansions of the
terms that comprise f :
sin(x) = x
x
3
3!
+ε(x)x
3
(sin(x))
2
=x
2
2x
4
3!
+ε(x)x
4
cos(x) =1
x
2
2!
+ε(x)x
3
x
2
cos(x) = x
2
+
x
4
2!
+ε(x)x
5
x
2
(1 cos(x)) =
x
4
2!
+ε(x)x
5
.
Now compute
lim
x0
sin
2
(x) x
2
cos(x)
x
2
(1 cos(x))
= lim
x0
1
6
x
4
+ε(x)x
4
x
4
2
+ε(x)x
5
=
1
6
+ε(x)
1
2
+ε(x) ·x
=
1
3
since ε(x)
x0
0.
55
Lemma 4.4.4. Let f : (a,b) 7→R be n +1 times differentiable. Suppose
f
0
(x) =b
0
+b
1
(x a) +b
2
(x a)
2
+···+b
n1
(x a)
n1
+ε(x)(x a)
n1
.
Let
g (x) = f (a)+b
0
(x a) +
b
1
2
(x a)
2
+···+
b
n1
n
(x a)
n
+
e
ε(x)(x a)
n
.
Then f (x) =g (x).
Proof. Compute g
0
(x) and realize that g
0
(x) = f
0
(x) and then use the result
that says if two derivatives are equal on an interval, then the functions are
equal up to a constant (Homework 9, Exercise 1) and we obtain
g (x) = f (x) +c
g (a) = f (a) +α
= α =0.
4.4.1 Local Extremum (Re-Visited)
Lemma 4.4.5. Let x
0
(a,b), f : (a, b) 7→ R be n times differentiable. As-
sume that f
(n)
is continuous at x
0
. Furthermore, assume that
f
0
(x
0
) = f
00
(x
0
) =···= f
(n1)
(x
0
) =0 and f
(n)
(x
0
) 6=0.
Then x
0
is a local minimum if f
(n)
(x
0
) >0 and a local maximum if f
(n)
<
0.
Proof (f
(n)
(x
0
) >0). Since f
(n)
(x
0
) > 0 and f
(n)
is continuous at x
0
, there is
δ >0 such that f
(n)
(x) >0 x (x
0
δ, x
0
+δ). By Taylors Theorem,
f (x) = f (x
0
) + f
0
(x
0
)(x x
0
)
| {z }
0
+···+
f
(n+1)
(x
0
)
(n 1)!
(x x
0
)
n1
| {z }
0
+
f
(n)
(c)
n!
(x x
0
)
n
| {z }
>0
= f (x) f (x
0
) x (x
0
δ, x
0
+δ)
which implies that x
0
is a local minimum by definition.
Example. Let f : (a,b) 7→ R be twice differentiable and x
0
(a,b) with f
00
(x
0
)
continuous and f
0
(x
0
) = 0. If f
00
(x
0
) > 0 then x
0
is a local minimum and if
f
00
(x
0
) <0 then x
0
is a local maximum.
56
For instance, consider f (x) = x
2
, f
0
(x) = 2x, f
00
(x) = 2, f
0
(0) = 0, f
00
(0) <
0 = 0 is a local maximum.
5 Integration
Definition 5.0.1. Let a < b. A partition σ of a closed interval (a,b) is a list of
ordered pairs such that
a = x
0
<x
1
<···<x
n
=b.
We write
σ
n
={x
0
,. .. , x
n
} =(x
i
)
n
i=1
.
Definition 5.0.2. The step of a partition σ
n
is defined as
h(σ
n
) = max
i=1,...,n
(x
i
x
i1
).
The particular case when
x
i
=x
i
+i ·
b a
n
i =0,...,n
is called the regular/uniform partition with
h(σ
n
) =
b a
n
.
Definition 5.0.3. Let f : [a,b] 7→R be continuous on [a,b] and σ
n
=(x
i
)
n
i=1
be
a partition of [a, b]. The real number
S
σ
(f ) =
n
X
i=1
m
i
(x
i
x
i1
)
where m
i
=min
x[x
i1
,x
i
]
f (x) is called the lower Darboux sum.
Similarly,
S
σ
(f ) =
n
X
i=1
M
i
(x
i
x
i1
)
where M
i
=max
x[x
i1
,x
i
]
f (x) is called the upper Darboux sum.
57
Remark. We have that
m(b a) S
σ
(f ) S
σ
(f ) M (b a)
where m and M are defined as above.
This implies that
S(f ) =inf{S
σ
(f ) : σ a partition of [a,b]}
S(f ) =sup{S
σ
(f ) : σ a partition of [a,b]}
are well-defined.
Lemma 5.0.4. Let σ,σ
be two partitions of [a,b] such that σ σ
. Then
S
σ
(f ) S
σ
(f ) and S
σ
S
σ
(f ).
Proof. Consider σ = {x
0
,. .. , x
n
} and σ
= {x
0
,. .. , x
n
}. Let x
j
i1
be the point in
σ
such that
x
i1
x
j
i1
x
i
j =0, ...,k
i
.
Then
S
σ
(f ) =
n
X
i=1
µ
min
x[x
i1
,x
i
]
f (x)
(x
i
x
i1
)
n
X
i=1
k
i
X
j =1
min
x[x
j
i1
,x
j
i
]
(x
j
i1
x
j 1
i1
)
=
n
X
i=1
min
x[x
i1
,x
i
]
f (x)(x
i
x
i1
) =S
σ
(f ).
Lemma 5.0.5. Suppose a < b and let f : [a,b] 7→ R be a continuous func-
tion. Then S(f ) =S(f ).
Proof. Since f is continuous on [a,b], f is uniformly continuous by Theorem
3.7.4. Recall that this means that ε >0 δ >0 such that
|x y|δ = |f (x) f (y)|
ε
3(b a)
.
Let σ be a partition with h(σ) δ and σ ={x
0
,. .. , x
n
}. Note that max
x[x
i1
,x
i
]
f (x) =
f (α
i
) α
i
[x
i1
, x
i
] by the Extreme Value Theorem (Theorem 3.6.14). Simi-
larly, lim
x[x
i1
,x
i
]
f (x) = f (β
i
) β
i
[x
i1
, x
i
]. This implies that
max
x[x
i1
,x
i
]
f (x) min
x[x
i1
,x
i
]
f (x) = f (α
i
) f (β
i
)
ε
3(b a)
58
since
|α
i
β
i
|δ
since α
i
,β
i
[x
i1
, x
i
] and h(σ) =max{x
i
x
i1
} δ.
Therefore,
S
σ
(f ) S
σ
(f ) =
n
X
i=1
(M
i
m
i
)(x
i1
x
i
)
ε
3
.
Additionally, by definition
S
σ
(f ) =sup{S
σ
(f ) : σ a partition of [a,b]}.
Therefore, by the characterization of suprema, there exists σ
1
, a partition of
[a,b], such that
S
σ
(f ) S
σ
1
ε
3
.
Similarly, there exists σ
2
, a partition of [a, b], such that
S
σ
2
(f ) S(f )
ε
3
.
Set σ =σ
1
σ
2
such that σ
1
σ and σ
2
σ (which is even smaller than σ
1
and
σ
2
, i.e., a refinement of the two). Using the result from the previous lemma,
we have
S
σ
(f ) S
σ
(f ) and S
σ
(f ) S
σ
1
(f )
which implies that
S
σ
(f ) S(f )
ε
3
and S(f ) S
σ
(f )
ε
3
.
Without loss of generality, we can assume that h(σ) δ (if it were not, we can
simply consider the uniform partition such that
ba
/n δ and perform the
same "trick"). Now, for every ε >0, we have
S(f ) S(f ) =S(f ) S
σ
(f )
| {z }
ε
3
+S
σ
(f ) S
σ
(f ) +S
σ
(f ) S(f )
| {z }
ε
3
= S(f ) =S( f ).
Definition 5.0.6 (Integral). Let a < b and f : [a,b] 7→ R be a continuous func-
tion. We call S( f ) =S(f ) the integral of f over [a,b]. We use the notation
Z
b
a
f (x) dx =S(f ) =S(f ).
59
Extensions:
If b < a, then
Z
b
a
f (x) dx =
Z
b
a
f (x)d x.
If b = a, then
Z
b
a
f (x)d x =0.
Remark. The integral
R
b
a
f (x)d x is really just notation for the sum
n
X
i=1
M
i
(x
i
x
i1
).
Lemma 5.0.7. Let a < b, f : [a,b] 7→ R be a continuous function. Further,
suppose (σ
n
)
n=1
be a sequence of partitions of [a,b] such that h(σ
n
)
n→∞
0.
Then
lim
n→∞
S
σ
n
(f ) =S(f ) and lim
n→∞
S
σ
n
(f ) =S(f )
where of course
S(f ) =S( f ) =
Z
b
a
f (x)d x.
Proof. Since f : [a,b] 7→R is continuous, f is uniformly continuous and ε >
0 δ >0,h(σ
n
) δ such that
S
σ
n
S
σ
n
ε (*)
n
X
i=1
(M
i
m
i
| {z }
ε
)(x
i
x
i1
) ε.
Note that because h(σ
n
)
n→∞
0, there exists N >0 such that h(σ
n
) δ n N.
In addition,
S
σ
n
S(f )
=
Z
b
a
f (x)d x
=S(f ) S
σ
n
(f ).
This means that n N,
0
S
σ
n
(f ) S(f ) ε
and
0 S
σ
n
(f ) S(f ) ε
60
i.e.,
lim
n→∞
S
σ
n
(f ) =S(f ) and lim
n→∞
S
σ
n
(f ) =S(f ).
Remark. The proof of the previous lemma can also be used to show that if
e
S
σ
n
(f ) =
n
X
i=1
f (
e
x
i
)(x
i
x
i1
)
e
x
i
[x
i1
, x
i
]
and h(σ
n
)
n→∞
0, then
lim
n→∞
e
S
σ
n
(f ) =
Z
b
a
f (x)d x.
Example. Let f (x) =e
x
, x [a,b]. Show that
Z
b
a
e
x
d x =e
b
e
a
.
First we want to compute S( f ). Consider
σ
n
=
x
0
=
a
, a +
b a
n
, a +2 ·
b a
n
,. .. , a +i ·
b a
n
,. .. , a +n ·
b a
n
=
b
.
Clearly, h(σ
n
) =
ba
/n and in particular h(σ
n
)
n→∞
0. Then compute
S
σ
n
(f ) =
n
X
i=1
m
i
(x
i
x
i1
)
=
ba
n
=
b a
n
·
n
X
i=1
e
a+(i 1)·
ba
n
=
b a
n
·e
a
·
n
X
i=1
³
e
ba
n
´
i1
=
b a
n
·e
a
·
n1
X
i=0
³
e
ba
n
´
i
=
b a
n
·e
a
Ã
1 e
ba
n
·n
1 e
ba
n
!
=e
a
(1 e
ba
) ·
ba
n
1 e
ba
n
.
61
where m
i
=min
x[x
i1
,x
i
]
e
x
=e
x
i1
. Now consider
lim
n→∞
Ã
e
a
(1 e
ba
) ·
ba
n
1 e
ba
n
!
= lim
x0
+
x
1 e
x
= lim
x0
+
1
e
x
=1.
Therefore, lim
n→∞
S
σ
n
(f ) =(e
ba
1)e
a
=e
b
e
a
=
R
b
a
f (x)d x.
5.1 Extension of Integrability
If σ is a partition of f bounded on [a,b] (not necessarily continuous), we can
define
S
σ
(f ) =
n
X
i=1
f
M
i
(x
i
x
i1
)
where
f
M
i
=sup
x[x
i1
,x
i
]
f (x) and similarly
S
σ
(f ) =
n
X
i=1
f
m
i
(x
i
x
i1
)
where
f
m
i
=inf
x[x
i1
,x
i
]
f (x).
Because f is bounded, we have m f (x) M x [a,b] m,M R which
subsequently means
m(b a) S
σ
(f ) S
σ
(f ) M (b a).
Then we have that the following are well-defined:
S(f ) =inf{S
σ
(f ) : σ a partition of [a,b]}
S(f ) =sup{S
σ
(f ) : σ a partition of [a,b]}.
However, note that S(f ) 6=S(f ) in general.
Definition 5.1.1 (Riemann integrability). Let f : [a, b] 7→R be a bounded func-
tion. We say f is Riemann integrable when S(f ) =S( f ).
5.2 Properties of Integrals
Lemma 5.2.1. Let a < b and f : [a,b] 7→R be a continuous function. Fur-
ther, let c [a,b]. Then
Z
b
a
f (x)d x =
Z
c
a
f (x)d x +
Z
b
c
f (x)d x.
62
Proof. Assume that c 6= a 6=b. Then consider
α
n
=
n
a, a +
c a
n
, a +2 ·
c a
n
,. .. ,c
o
={x
i
}
n
i=0
β =
½
c,c +
b c
n
,c +2 ·
b c
n
,. .. ,b
¾
={x
i
}
2n+1
i=n
where both α
n
and β
n
are regular partitions of [a,c]/[c,b], respectively. Then
consider
σ
n
=α
n
β
n
,
which is a regular partition of [a,b] with h(σ
n
) = max{h(α
n
),h(β
n
)}
n→∞
0.
Then compute
Z
b
a
f (x)d x = lim
n→∞
S
σ
n
(f ) = lim
n→∞
Ã
n
X
i=1
M
i
(x
i
x
i1
+
2n+1
X
i=n+1
M
i
(x
i
x
i1
)
!
= lim
n→∞
³
S
σ
n
(f ) +S
σ
n
(f )
´
=
Z
c
a
f (x)d x +
Z
b
c
f (x)d x.
Lemma 5.2.2. Let f : [a,b] 7→ R be a continuous function and let λ R.
Then
Z
b
a
f (x)d x =λ
Z
b
a
f (x)d x.
Proof. Let {σ
n
}
nN
be a sequence of partitions with h(σ
n
)
n→∞
0. Then we
have
Z
b
a
(λf )(x)d x = lim
n→∞
³
S
σ
n
(λf )
´
= lim
n→∞
n
X
i=1
max
x[x
i1
,x
i
]
(λf )(x)(x
i
x
i1
)
=λ · lim
n→∞
n
X
i=1
max
x[x
i1
,x
i
]
f (x)
| {z }
S
σ
n
(f )
(x
i
x
i1
)
=λ
Z
b
a
f (x)d x.
63
Lemma 5.2.3 (Integral Mean Value Theorem). Let f : [a,b] 7→R be a con-
tinuous function. Then there exists c [a,b] such that
f (c) =
1
b a
Z
b
a
f (x)d x.
Proof. Since f is continuous on [a,b], the Extreme Value Theorem (Theorem
3.6.14) implies that
m = min
x[a,b]
f (x) f (x) max
x[a,b]
=M .
Let f (α) =min
x[a,b]
f (x) and f (β) =max
x[a,b]
. Then we have
m(b a)
Z
b
a
f (x)d x M (b a)
m
1
b a
Z
b
a
f (x)d x M .
Then, by the Intermediate Value Theorem (Theorem 3.6.9), there exists some
c [a,b] such that
f (c) =
1
b a
Z
b
a
f (x)d x.
5.3 Antiderivatives
Definition 5.3.1 (Antiderivative). Let f : [a,b] 7→R be a continuous function.
We say F : [a,b] 7→R is an antiderivative of f on [a,b] if:
1. F is continuous on [a,b] and
2. F
0
(x) = f (x) x (a,b).
Remark. If F and G are two antiderivatives of f on [a,b], then F
0
= f and
G
0
= f implies that F
0
=G
0
on (a,b).
Lemma 5.3.2. Let f : [a,b] 7→R be a continuous function. Then
F (x) =
Z
x
a
f (t)d t = lim
n→∞
S
σ
n
(x)
(f )
is an antiderivative of f on [a,b].
64
Proof. We first show that F
0
(x
0
) = f (x
0
), x
0
(a,b). Let x
0
(a,b),x (a,b) \
{x
0
}. The IMVT (Theorem 5.2.3) on the interval (x
0
, x) or (x, x
0
) implies that
there exists some c(x) [x
0
, x] such that
f (c(x))(x x
0
) =
Z
x
x
0
f (t)d t.
We have
Z
x
x
0
f (t)d t =
Z
x
a
f (t)d t
Z
x
0
a
f (t)d t
=F (x) F (x
0
)
f (c(x)) =
F (x) F (x
0
)
x x
0
lim
xx
0
f (c(x)) = f
µ
lim
xx
0
c(x)
= f (x
0
)
f (x
0
) = lim
xx
0
F (x) F (x
0
)
x x
0
= f
0
(x
0
)
and we have f (x) =F
0
(x), x (a,b) and F is continuous on (a,b).
It remains to show that
lim
xa
+
F (x) =F (a) and lim
xb
F (x) =F (b).
Take x [a,b]. By the IMVT, there exists some c(x) [a,x] such that f (c(x))(x
a) =
R
x
a
f (t)d t =F (x) and
lim
xa
+
f (x) = lim
xa
+
f (c(x))(c a) =0 =F(a).
The argument is identical for F (b) = f (b).
Lemma 5.3.3. Let f : [a,b] 7→R be a continuous function. Further, let h,g :
(c,d) 7→[a,b] be differentiable on (c,d). Then
K (x) =
Z
g (x)
h(x)
f (t)d t
is differentiable on (c, d) and K
0
(x) = f (g (x))g
0
(x) f (h(x))h
0
(x) for some
x (c, d).
Proof. Let
e
f (x) =
f (x) x [a,b]
f (a) x [a 1, a]
f (b) x [b, b +1]
65
so
e
f is continuous on the interval [a 1, b +1]. Then we have
F (x) =
Z
x
a1
e
f (t)d t and K (x) =F (g (x))F(h(x))
with
K
0
(x) =F
0
(g (x))g
0
(x) F
0
(h(x))h
0
(x)
=
e
f (g (x))g
0
(x)
e
f (h(x))h
0
(x)
= f (g (x))g
0
(x) f (h(x))h
0
(x).
Theorem 5.3.4 (Fundamental Theorem of Calculus). Let f : [a, b] 7→R be
continuous and G be an antiderivative of f on [a,b]. Then
Z
b
a
f (t)d t =G(b)G(a) =G
¯
¯
¯
b
a
and
Z
b
a
G
0
(t )d t =G(b) G(a).
Proof. We know that
F (x) =
Z
x
a
f (t)d t
is an antiderivative of f which implies that F(x) =G(x)+c. Then we have that
F (b) =G(b)+c
F (a) =G(a)+c
and subsequently
Z
b
a
f (t)d t =G(b)G(a)
since F (a) is zero by definition.
5.4 Properties of Integrals
Lemma 5.4.1. Let f , g : [a,b] 7→ R be continuous functions and let λ R.
Then
Z
b
a
(λf +g )(x)dx =λ
Z
b
a
f (x)d x +
Z
b
a
g (x)dx.
66
Proof. Take
K (x) =
Z
x
a
(λf +g )(t)d t λ
Z
x
a
f (t)d t
Z
x
a
g (t )d t.
Then consider that K
0
(x) = (λf +g )(x) λ f (x) g(x) = 0 and K (x) = c c R
by the IMVT. Then we have K (b) =K (a) =0.
Lemma 5.4.2. Let f : [a,b] 7→R be a continuous function such that f (x)
0 x [a,b]. Then, for every c [a,b],
0
Z
c
a
f (x)d x
Z
b
a
f (x)d x
and furthermore
Z
b
a
f (x)d x =0 f (x) =0 x [a,b].
Proof. Left as an exercise.
5.5 Change of Variable
Lemma 5.5.1 (Change of Variable). Let f : [a, b] 7→R be a continuous func-
tion and I be an open interval. Further, let ϕ : I 7→ R be a continuous and
differentiable function on I (ϕ
0
is also continuous on I) such that such that
[α,β] I with ϕ([a,b]) [a,b]. Then
Z
ϕ(β)
ϕ(α)
f (x)d x =
Z
β
α
f (ϕ(t))ϕ
0
(t )d t.
We write x =ϕ(t ) is a change of variable.
Proof. Let G(x) =
R
ϕ(x)
0
f (t)d t and g (x) = f (ϕ(x))ϕ
0
(x). Both of these func-
tions are continuous on [α,β]. Furthermore, we have G
0
(x) =g (x) and G is an
antiderivative of g . So by the Fundamental Theorem of Calculus, we have
G(β) G(α) =
Z
β
α
g (t )d t
Z
ϕ(β)
0
f (x)d x
Z
ϕ(α)
0
f (x)d x =
Z
β
α
f (ϕ(t))ϕ
0
(t )d t
Z
ϕ(β)
ϕ(α)
f (x)d x =
Z
β
α
f (ϕ(t))ϕ
0
(t )d t.
67
Example. Compute
J =
Z
1
0
1
(x
2
+1)
3
2
.
In this context, we have f : [0,1] 7→ R, f (x) = 1/(x
2
+1)
3/2
. Take ϕ(t) = tan(t )
such that ϕ : (π/2,π/2) 7→R and ϕ
0
(t ) =1+tan
2
(t ). Apply change of variables:
J =
Z
1
0
1
(x
2
+1)
3
2
=
Z
ϕ(0)=0
0
π
/4
ϕ(
π
/4)=1
1 +tan
2
(t )
(tan
2
(t ) +1)
3
/2
d t
=
Z
π
/4
0
1
(1 +tan
2
(t )
1
/2
)
d t
=
Z
π
/4
0
1
³
1
cos(t )
´
1
/2
d t
=
Z
π
/4
0
cos(t )d t =sin(t )
¯
¯
¯
π
/4
0
=
p
2
2
.
Remark. The function ϕ does not need to be invertible.
5.6 Integration by Parts
Lemma 5.6.1 (Integration by Parts). Let f , g : [a,b] 7→R be two continuous
and differentiable functions with both f
0
and g
0
continuous on [a,b]. Then
Z
b
a
f
0
(x)g (x)dx =
Z
b
a
f (x)g
0
(x)dx + f (x)g (x)
¯
¯
¯
b
a
.
Proof. Note that
(f ·g )
0
= f
0
·g + f ·g
0
.
Thus, f ·g is an antiderivative of f
0
·g + f ·g
0
. Therefore,
f (b)g (b) f (a)g (a) =
Z
b
a
¡
f
0
(x)g (x) + f (x)g
0
(x)
¢
d x.
So, integration by parts is nothing more than the Fundamental Theorem of
Calculus applied to a product of functions.
Examples.
68
1. Compute
J(x) =
Z
x
0
e
t
=
g
0
(t )
=
g (t)=e
t
cos(t )
=
f (t)
d t =
Z
x
0
e
t
=
g
0
(t )
(sin(t ))
=
f (t)
d t +cos(x) ·e
x
cos(0) ·e
0
| {z }
1
=
Z
x
0
e
t
·cos(t )
| {z }
J(x)
+sin(x) ·e
x
sin(0) ·e
0
+cos(x) ·e
x
1
= 2J(x) =sin(x) ·e
x
+cos(x) ·e
x
1
J(x) =
1
2
(sin(x) +cos(x))·e
x
1
2
.
2. Take n 1. Compute
J
n
(x) =
Z
x
0
1
(t
2
+1)
n
d t.
Take g (t ) =1/(t
2
+1)
n
and f (t) =t such that g
0
(t ) =n(t
2
+1)
(n+1)
and
f
0
(t ) =1. Then we have
J
n
(x) =
Z
x
0
t(2t n)
(t
2
+1)
n+1
d t +
x
(x
2
+1)
n
0
=2n
Z
x
0
t
2
+1
(t
2
+1)
n+1
d t 2n
Z
x
0
1
(t
2
+1)
n+1
d t +
x
(x
2
+1)
n
=2n ·J
n
(x) 2n ·J
n+1
(x) +
x
(x
2
+1)
n
J
n+1
(x) =
2n 1
2n
·J
n
(x) +
1
2n
·
x
(x
2
+1)
n
.
5.7 Taylor Expansion
Lemma 5.7.1. Let f : [a,b] 7→ R be (n +1) times differentiable and f
(n+1)
be continuous on [a,b]. Then
f (x) = f (a)+f
0
(a)(x a)+···+
f
(n)
(a)
n!
(x a)
n
+
1
n!
Z
x
a
(x t)
n
f
(n+1)
(t )d t.
Proof. Using integration by parts, compute for k N,n 1
Z
x
a
(x t )
k1
| {z }
u
0
(t )
f
(k)
(t )
| {z }
v(t)
d t =
1
k
Z
x
a
(x t )
k
f
(k+1)
(t )d t +
µ
1
k
(x t )
k
f
(k)
(t )
¯
¯
¯
x
a
=
1
k
Z
x
a
(x t )
k
f
(k+1)
(t )d t +
1
k
(x a)
k
f
(k)
(a).
69
We now proceed by induction on n. For n = 0, we can simply use the Funda-
mental Theorem of Calculus to obtain
f (x) f (a) =
Z
x
a
f
0
(t )d t
and P(0) is true. Now, assume that P(k 1) is true, i.e.,
f (x) = f (a)+···+
f
(k1)
(a)
(k 1)!
(x a)
k1
+
1
(k 1)!
Z
x
a
(x t )
k1
f
(k)
(t )d t.
Using the formula we just proved, we have that
f (x) = f (a)+···+
f
(k1)
(a)
k!
(x a)
k1
+
1
(k 1)!
µ
1
k
(x a)
k
f
(k)
(a)
+
1
k
Z
x
a
(x t )
k
f
(k+1)
(t )d t
which implies that P(k) is true.
Example. Apply this for n =1 and obtain
f (x) = f (a)+ f
0
(a)(x a) +
Z
x
a
(x a)f
00
(t )d t
f
0
(a) =
f (x) f (a)
x a
1
x a
Z
x
a
(x t )f
00
(t )d x
| {z }
error"
¯
¯
¯
¯
f
0
(a)
f (x) f (a)
x a
¯
¯
¯
¯
1
x a
Z
x
a
|(x t )|·|f
00
(t )|d t sup
[a,b]
¯
¯
f
00
(t )
¯
¯
1
2(x a)
(x a)
2
sup
[a,b]
¯
¯
¯
f
00
(t ) ·
x a
2
¯
¯
¯
.
Take the ODE:
y
0
(t ) = f (t , y(t)), y(0) =α,t [0,1].
Let σ
n
be a regular partition of [0,1] such that h(σ
n
) =
1
/n. Then we have that
y
0
(x
i
)
y(x
i+1
) y(x
i
)
h
=
x
i+1
x
i
and
¯
¯
¯
¯
y
0
(x
i
)
y(x
i+1
) y(x
i
)
h
¯
¯
¯
¯
sup
[a,b]
¯
¯
y
00
(x)
¯
¯
·
h
2
.
This is a valid solution to the ODE since
y(x
i+1
) y(x
i
)
h
f (x
i
, y(x
i
)).
Let Y
i
be an approximation of y(x
i
) defined by:
Y
i
=
(
Y
i+1
Y
i
h
= f (x
i
,Y
i
) i =0,1,...,n 1
Y
0
=α.
70
This is equivalent to
Y
i
=
(
Y
i+1
=Y
i
+h · f (x
i
,Y
i
) i =0,1,...,n 1
Y
0
=α.
5.8 Weak Derivatives
Definition 5.8.1 (Weak derivative). Let f , g : [a,b] 7→ R be two continuous
functions with
Z
b
a
f
0
(x)g (x)dx =
Z
b
a
f (x)g
0
(x)dx + f g
¯
¯
¯
b
a
.
Now assume that g is smooth and g (a) = g (b) = 0. With this condition, we
now have
Z
b
a
f
0
(x)g (x)dx =
Z
b
a
f (x)g
0
(x)dx.
We say that p : [a,b] 7→ R is the weak derivative of f provided that p is inte-
grable and
R
b
a
pg =
R
b
a
f g
0
for every g smooth with g (c) =g (b) =0.
Remark. If p exists, it is unique and if f is differentiable, then p = f
0
.
Example. Take f (x) =|x|, x [1, 1]. We have
Z
1
1
=
Z
1
1
|x|·g
0
(x)dx g smooth ,g(1) =g (1) =0.
But
Z
1
1
|x|·g
0
(x)dx =
Z
0
1
x ·g
0
(x)
Z
1
0
x ·g
0
(x)
=
Z
0
1
1 ·g (x) +
x ·g (x)
¯
¯
¯
0
1
+
Z
1
0
g (x)dx +
x ·g (x)
¯
¯
¯
1
0
(IBP)
=
Z
1
1
p(x)g (x)d x
where
p(x) =
(
1 x < 0
1 x 0.
Definition 5.8.2 (Diraq measure). Consider the function
δ
n
(x) =
0 x >
1
n
0 x <
1
n
n
2
1
n
x
1
n
.
71
Compute
Z
1
1
f (x)δ
n
(x)dx =
Z
1
n
1
n
f (x)·
n
2
d x =
1
2
Z
1
n
1
n
n·f (x)
n→∞
f (0) =
Z
1
1
f (x)δ
0
(x)dx.
This is called the Dirac measure.
Example. Now take
f (x) =
(
1 x 0
0 x < 0.
Now try to make sense of a weak derivative. Consider g smooth such that
g (1) =g (1) =0. Then we have
Z
1
1
f (x)g
0
(x)dx =
Z
0
1
g
0
(x)dx
=g (1) +g(0) (FTC)
=g (0) =
Z
1
1
δ
0
(x)g (x)dx.
72
List of Theorems
1.1.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Definition (Boundedness) . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.2 Definition (Infimum) . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.3 Definition (Supremum) . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.4 Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.5 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.6 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.7 Definition (Absolute value) . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.8 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.1 Definition (Sequential convergence) . . . . . . . . . . . . . . . . . . 8
2.1.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.3 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.4 Definition (Sequential monotonicity) . . . . . . . . . . . . . . . . . 10
2.1.5 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3.2 Theorem (Squeeze Theorem) . . . . . . . . . . . . . . . . . . . . . . 13
2.4.1 Definition (limsup and liminf) . . . . . . . . . . . . . . . . . . . . . 14
2.4.2 Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.1 Definition (Subsequences) . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.3 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.4 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6.1 Definition (Peak points) . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6.2 Lemma (Monotone Subsequence) . . . . . . . . . . . . . . . . . . . 16
2.6.3 Theorem (Bolzano-Weirstrass) . . . . . . . . . . . . . . . . . . . . . . 17
2.6.4 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.7.1 Definition (Cauchy sequences) . . . . . . . . . . . . . . . . . . . . . 19
2.7.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.8.1 Definition (Divergence to ±∞) . . . . . . . . . . . . . . . . . . . . . 21
2.8.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.0.1 Definition (Functional limits) . . . . . . . . . . . . . . . . . . . . . . 22
3.0.2 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.1 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1.2 Lemma (Cauchy Criteria for Functions) . . . . . . . . . . . . . . . . 25
3.2.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.3 Theorem (Squeeze Theorem) . . . . . . . . . . . . . . . . . . . . . . 27
3.4.1 Lemma (Composition of Functions) . . . . . . . . . . . . . . . . . . 28
3.5.1 Definition (Infinite functional limits) . . . . . . . . . . . . . . . . . . 28
3.5.2 Definition (Infinite functional limits) . . . . . . . . . . . . . . . . . . 29
3.5.3 Definition (Combined infinite functional limits) . . . . . . . . . . . 29
73
3.6.1 Definition (Functional continuity) . . . . . . . . . . . . . . . . . . . 29
3.6.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.6.3 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.6.4 Definition (Continuity on an interval) . . . . . . . . . . . . . . . . . 31
3.6.5 Definition (Left/right limits) . . . . . . . . . . . . . . . . . . . . . . . 31
3.6.6 Definition (Left/right continuous) . . . . . . . . . . . . . . . . . . . 31
3.6.7 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.6.8 Definition (Continuity on a closed interval) . . . . . . . . . . . . . . 32
3.6.9 Theorem (Intermediate Value Theorem) . . . . . . . . . . . . . . . . 32
3.6.10 Definition (Monotone functions) . . . . . . . . . . . . . . . . . . . . 34
3.6.11 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.6.12 Definition (Boundedness on a set) . . . . . . . . . . . . . . . . . . . 34
3.6.13 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.6.14 Theorem (Extreme Value Theorem) . . . . . . . . . . . . . . . . . . . 35
3.6.15 Definition (Supremum and max of a function) . . . . . . . . . . . . 36
3.7.1 Definition (Uniform continuity) . . . . . . . . . . . . . . . . . . . . . 36
3.7.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.7.3 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.7.4 Theorem (Uniform continuity) . . . . . . . . . . . . . . . . . . . . . 38
3.7.5 Theorem (Continuity Extension) . . . . . . . . . . . . . . . . . . . . 39
3.8.1 Proposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.8.2 Theorem (Inverse continuity) . . . . . . . . . . . . . . . . . . . . . . 42
4.0.1 Definition (Locally differentiable) . . . . . . . . . . . . . . . . . . . . 42
4.0.2 Definition (Globally differentiable) . . . . . . . . . . . . . . . . . . . 43
4.0.3 Theorem (Caratheodory’s Theorem) . . . . . . . . . . . . . . . . . . 44
4.0.4 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.1.1 Definition (Local extremum) . . . . . . . . . . . . . . . . . . . . . . . 45
4.1.2 Theorem (Fermat) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.1 Proposition (Basic rules of differentiation) . . . . . . . . . . . . . . . 46
4.2.2 Proposition (Product Rule) . . . . . . . . . . . . . . . . . . . . . . . . 46
4.2.3 Proposition (Quotient Rule) . . . . . . . . . . . . . . . . . . . . . . . 47
4.2.4 Proposition (Chain Rule) . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3.1 Theorem (Rolles Theorem) . . . . . . . . . . . . . . . . . . . . . . . . 49
4.3.2 Theorem (Mean Value Theorem) . . . . . . . . . . . . . . . . . . . . 49
4.3.3 Theorem (Mean Value Theorem (Cauchy)) . . . . . . . . . . . . . . 50
4.3.4 Proposition (L’Hospital’s Rule) . . . . . . . . . . . . . . . . . . . . . . 51
4.4.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4.2 Theorem (Taylor) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4.3 Lemma (Taylor Expansion uniqueness) . . . . . . . . . . . . . . . . 53
4.4.4 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.4.5 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.0.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.0.2 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.0.3 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.0.4 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.0.5 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
74
5.0.6 Definition (Integral) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.0.7 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.1.1 Definition (Riemann integrability) . . . . . . . . . . . . . . . . . . . 62
5.2.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.2.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2.3 Lemma (Integral Mean Value Theorem) . . . . . . . . . . . . . . . . 64
5.3.1 Definition (Antiderivative) . . . . . . . . . . . . . . . . . . . . . . . . 64
5.3.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.3.3 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.3.4 Theorem (Fundamental Theorem of Calculus) . . . . . . . . . . . . 66
5.4.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.4.2 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.5.1 Lemma (Change of Variable) . . . . . . . . . . . . . . . . . . . . . . . 67
5.6.1 Lemma (Integration by Parts) . . . . . . . . . . . . . . . . . . . . . . 68
5.7.1 Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.8.1 Definition (Weak derivative) . . . . . . . . . . . . . . . . . . . . . . . 71
5.8.2 Definition (Diraq measure) . . . . . . . . . . . . . . . . . . . . . . . . 71
75