Lecture Notes AM1
Lecture Notes AM1
Institut für Algebra und Geometrie, Karlsruher Institut für Technologie, October
2023
3
Introduction
These lecture notes are companions of my lecture on this topic in the winter term
2021/22. Although many things in the lectures were inspired by the lecture notes
of Prof. Axenovich, I often had another focus. Still the lecture notes owe quite a
bit to its anchestor.
We will treat the basic objects of (real) analysis. Based on studying sequences
and series we will study the notions of continuity, differentiability and inegration.
We will then consider first examples of differential equations.
It takes some patience until we will be at the point where you want to be, but in
order to be able to apply mathematics properly it is helpful to understand WHY
things are as stated. Therefore we will prove several statements quite thoroughly,
because they should be true always. This will involve certain proofs where you
will think there is nothing to prove. This might then be a consequence that
in mathematics things are defined in a strict way even if we have an intuitive
understanding, and that we have to make shure that our definitions do what we
expect them to do.
I do not include many pictures, because they are part of the lectures. The pictures
I did include were created by me. Try to find the right pictures for the things
we write down. This would be a good starting point to understand the abstract
world of mathematics.
3 Sequences 39
3.1 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 Subsequences and accumulation points . . . . . . . . . . . . . . . 46
4 Continuity 55
4.1 Basic definitions and examples . . . . . . . . . . . . . . . . . . . . 55
4.2 Two Important Theorems . . . . . . . . . . . . . . . . . . . . . . 57
5 Series 63
5.1 Infinite sums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Criteria for absolute convergence . . . . . . . . . . . . . . . . . . 70
5.3 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.4 The exponential function and some relatives . . . . . . . . . . . . 74
5.5 Two examples for calculations with power series . . . . . . . . . . 80
6 Differentiability 83
5
6 INHALTSVERZEICHNIS
7 Integration 105
7.1 Lebesgue Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.2 The main theorems on integration . . . . . . . . . . . . . . . . . . 111
7.3 Calculating some integrals . . . . . . . . . . . . . . . . . . . . . . 116
7.4 Some elementary differential equations . . . . . . . . . . . . . . . 120
7.5 Improper integrals . . . . . . . . . . . . . . . . . . . . . . . . . . 127
7.6 Integrals of rational functions . . . . . . . . . . . . . . . . . . . . 131
8 Indices 135
8.1 Important Theorems . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.2 Some Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.3 Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Kapitel 1
1.1 Sets
The notion of sets and constructions from set theory are basic tools in all of
mathematics. We nevertheless have to start with a naive definition which would
have to be stated differently if we really would want to spend time on the basics
of mathematics. This would take too long, and for our needs the naive version
will do.
Remark 1.1.2 There are no restrictions what the objects of a set can be. They
can be rather complicated, in particular the elements of a set can themselves
can be sets as well. Everything can be an object in a set. We will very often be
concerned with more modest sets, for instance sets of numbers.
Things might become more clear by looking at examples (see next number).
A set can be written in several ways. Sometimes we can write up all elements, as
in S = {2, 3, 5, 7}. Sometimes it is more clever to write down properties which
correspond to being an element of a specific set. We then mark this condition
by a vertical line or by a colon. For instance, the above set with four elements
can also be written as S = {p | p is a prime number and p ≤ 10}. In particular,
infinite sets often need to be described in that way.
7
8 KAPITEL 1. BASIC NOTIONS FROM MATHEMATICS
Example 1.1.3 The following are examples for sets which you should have seen
before:
• N, the set of all natural numbers, its elements are 1, 2, 3, . . . and we then
write N = {1, 2, 3, . . . }, hoping that everyone agrees what the dots should
mean. If we want to include 0 to this set, we call it N0 .
• Z, the set of all integers, consisting of all natural numbers, zero, and all
elements −n for n ∈ N.
Z = {x | x ∈ N0 or − x ∈ N0 }.
S ∩ ∅ = ∅, S ∪ ∅ = S.
S ∩ T = T ∩ S, S ∪ T = T ∪ S.
(S ∪ T ) ∪ U = S ∪ (T ∪ U ), (S ∩ T ) ∩ U = S ∩ (T ∩ U ),
S ∩ (T ∪ U ) = (S ∩ T ) ∪ (S ∩ U ), S ∪ (T ∩ U ) = (S ∪ T ) ∩ (S ∪ U ).
The identities in the first line are quite clear: There is no element in ∅, and
therefore also no element in ∅ which belongs to S. The elements which belong
to S or to ∅ are exactly the elements in S.
The second and third line are left as an exercise.
The first identity in the fourth line means: An element belongs to S and T or U
if and only if it belongs to S and T or to S and U. Written more formally this
looks like: Let x be an object. If x belongs to S and to T ∪ U then it belongs to
T or U. If it belongs to T then it belongs to S ∩ T hence to (S ∩ T ) ∪ (S ∩ U )
and similarly if it belongs to U. Therefore
S ∩ (T ∪ U ) ⊆ (S ∩ T ) ∪ (S ∩ U ).
S r T := {s ∈ S | s 6∈ T },
S × T = {(s, t) | s ∈ S and t ∈ T }.
10 KAPITEL 1. BASIC NOTIONS FROM MATHEMATICS
This is the product of S and T , its elements are called pairs. If S = T, we write
S 2 instead of S × S.
We similarly define
S × T × U = {(s, t, u) | s ∈ S, t ∈ T, u ∈ U }
which has the property that for every d ∈ D there exists a unique c ∈ C (namely
c = f (d) ) such that (d, c) ∈ G(f ).
If, to the contrary, we are given a set G ⊆ D × C such that for every d ∈ D
there exists exactly one c ∈ C such that (d, c) ∈ G, this set G is the graph of a
map.
If the codomain of a map is contained in R we will often call f a function. As
soon as we have introduced complex numbers, also maps with complex values are
functions. We will not be too strict with this wording.
1.1. SETS 11
The maps g ◦ f and f ◦ g are “the identity” on D resp. C, i.e. the maps which
map every element from their domain to itself.
Every set has its own identity map.
Sometimes one is interested in inverting functions. You know examples of this:
The square root function is inverse to the squaring function (both on non-negative
real numbers):
√
f : R≥0 → R≥0 , f (x) = x2 , g : R≥0 → R≥0 , g(y) = y
are inverses to one another.
And if you want to know which number x satisfies ex = y for a given positive
y, you have to invent the natural logarithm, which is invers to the exponential
map.
In order that a map is invertible (i.e. an inverse map exists) you need that every
element of C is realized as a value of f, i.e. the range is all of C (“ f is surjective”)
and that no value is assumed more often than once (“ f is injective”). We will not
make extensive use of this terminology, but sometimes it is helpful. In particular,
the squaring function is not injective on all of R (as, e.g., 12 = (−1)2 ) and is
not surjective if we use R as the codomain, because the square of a real number
never is negative.
If D, C ⊆ R and f : D → C is invertible, the graph of the inverse function is
produced from the graph of f by reflecting it along the diagonal in the first/third
quadrant.
1.2 Inequalities
For any two real numbers x, y we say that x is larger than y if x − y is positive.
In symbols:
x > y iff x − y > 0.
Here, “iff” is an abbreviation for “if and only if” which is quite common in
mathematical texts.
1.2. INEQUALITIES 13
This is the transitivity of the order relation. The additive monotonicity of the
order relation says that for any x, y, z ∈ R we have
x < y ⇒ x + z < y + z,
and the multiplicative monotonicity says that for all x, y, z ∈ R with z > 0 we
have
x < y ⇒ xz < yz.
These rules are well known from highschool.
We also denote by x ≤ y that x < y or x = y, and similarly x ≥ y means
x > y or x = y. Both ≤ and ≥ satisfy transitivity, additive mononicity and
multiplicative monotonicity but not the trichotomy (because ≤ and = are not
exclusive). The replacement for trichotomy is
[x ≤ y and y ≤ x] ⇒ x = y.
The absolute value satisfies an utmost important rule which will be used again
and again in many situations:
1
If A, B are two assertions, we write A ⇒ B if the truth of A implies the truth of B.
14 KAPITEL 1. BASIC NOTIONS FROM MATHEMATICS
|x + y| ≤ |x| + |y|.
This is clear if x and y are either both positive or both negative, and if one of
them is zero. Then we even have equality. The interesting case is x > 0, y < 0
(or vice versa, which goes similarly). Then if x + y ≥ 0, we have
We have to take care that this is not mistaken with the pair having a and b as
its components. The context (or an explicite remark) should make clear, which
meaning the notation has in the specific situation. Note that (a, a) = ∅.
We also have half open intervals
[a, ∞) := {x ∈ R | x ≥ a}
and similarly (a, ∞), (−∞, b), (−∞, b], but we do not include ∞ in the interval,
as this fails to be a real number.
The last special case is (−∞, ∞) = R.
1.3. METHODS OF PROOF 15
x ≤ y ⇒ f (x) ≤ f (y).
(e.g. what is a set, what are numbers, what does addition mean?) and which in
pure“ mathematics is reduced to a smallest possible list of truths which are
”
accepted. In this course we take a larger base coming from your experience in
school, although we will prove many facts which you already know from school.
Given this base of true statements we will derive everything else by that using
rigid arguments which then are called proofs. There are several schemes of argu-
mentation which may be used for proving new statements. You will see proofs
over and over again, but it might be helpful to describe two general ways to argue
and one specific way to deal with natural numbers.
As ã2 and b̃2 are odd, after cancelling 2max(2f +1, 2e) we are left with the equality
of an even and an odd number, which gives the desired contradicton.
1.3. METHODS OF PROOF 17
s=x−q
0.9999999999 . . . = 0.9 = 1.
The second equation is the one in question, the first one only illustrates what
the notation 0.9 means: a periodic decimal number. If this number 0.9 were
different from 1, it certainly were the largest real number smaller than 1. But
such a number does not exist, because for any two different real numbers x < y,
(in particular for y = 1 ) their arithmetic mean x+y
2
satisfies
x+y
x< < y.
2
18 KAPITEL 1. BASIC NOTIONS FROM MATHEMATICS
S := {n ∈ N | A(n) is false}
is empty.
Assume that S is non-empty. Then – as indicated in 1.2.6 – S contains a smallest
element. This shows that if we can demonstrate that S does not contain a smallest
element we know it has to be empty. This idea creates what is called mathematical
induction.
In order to show that A(n) is true for every n ∈ N, it suffices to show that A(1)
is true (the base case) and that the truth of A(n) for any single natural number
n implies the truth of A(n + 1) (the induction step).
Note that we do not assume that A(n) is true for every n in order to prove
A(n + 1) , but that we only assume that A(n) s true for one (arbitrary) n.
In other words we show every of the implications
and as A(1) is true by the base case, we successively see that A(2) also has to
be true, hence also A(3) , hence A(4) , hence. . . .
An example should make this procedure more clear.
(1 + h)n ≥ 1 + nh.
(1 + h)n+1 ≥ 1 + (n + 1)h.
Here we use that h ≥ −1, as this implies 1 + h ≥ 0 , hence we can use the
multiplicative monotonicity of ≥ (cf. 1.2.1) in (1) , and as nh2 ≥ 0, we can use
the additive monotonicity of ≥ in step (2).
(1 + h)2 = 1 + 2h + h2 ≥ 1 + 2h.
It also holds for h ≥ −2 and every n but the proof by induction no longer works
for −2 ≤ h < −1, because 1 + h < 0 and therefore montonicity of ≥ cannot be
applied in this case.
However, for these values of h, |1+h| ≤ 1 and therefore |(1+h)n | = |1+h|n ≤ 1,
hence (1 + h)n ≥ −1, while 1 + nh < −1 for n ≥ 3.
We will soon have many occasions to see other proofs by mathematical induction.
20 KAPITEL 1. BASIC NOTIONS FROM MATHEMATICS
Kapitel 2
21
22 KAPITEL 2. NUMBERS AND FUNCTIONS
as desired. Here, (IH) is the inductive hypothesis, i.e the assertion for the
value n under consideration.
More generally, for given values a, b ∈ R, the numbers
ai := b + ia, i = 0, 1, 2, 3, . . .
as all other powers of q occur in the sum with positive and with negati-
ve sign as a summand, and these two summands cancel, leaving only the
highest and lowest power at the end.
A proof by mathematical induction also is possible.
2.1. SUMS AND PRODUCTS 23
(c) In a similar way as the first example, we can show for instance that
n
X 1 1
i2 = · n · (n + ) · (n + 1).
i=0
3 2
The letter Π is the capital “pi”, the greek counterpart to our “p”, the initial of
the word “product”. Sometimes the multiplication index i can also go through
an other finite set of integers (or even more general - see later) between a lower
and upper bound, for instance
n
Y
ai , if m ≤ n.
i=m
We again
Q use a similar notation for a finite set S and a function a defined on
S: s∈S a(s). If S = U ∪ V, where U ∩ V = ∅, we again have
Y Y Y
a(s) = a(u) · a(v).
s∈S u∈U v∈V
The number n! is the number of different ways to order n different items. For
instance, there are 3! = 6 ways to write down the numbers 1, 2, 3 in different
orders:
123, 132, 213, 231, 312, 321.
We have (n+1)! = (n+1)·n! for every n. More generally, we have for 0 ≤ k ≤ n
n
Y
n! = k! · i.
i=k+1
Cancelling the factor k! in numerator and denominator, this can also be written
as Qn
n i
= i=k+1 .
k (n − k)!
These numbers are called the binomial coefficents.
The combinatorial meaning of the binomial coefficients is that nk is the number
Proof. There are several ways to prove this theorem. We show two of them, you
can choose your favourite one.
2.1. SUMS AND PRODUCTS 25
(k+1)·n!+(n−k)·n!
= (k+1)!·(n−k)!
(n+1)! n+1
= (k+1)!·(n+1−(k+1))!
= k+1
.
From the first to the second line we have expanded the fractions to the
common denominator (k + 1)! · (n − k)! and then added the numerators.
1
1 1
1 2 1
1 3 3 1
1 4 6 4 1
1 5 10 10 5 1
and so on. Every entry in the next row is the sum of the two entries left and right
of it in the row before. The missing entries left and right of the 1s are zero.
Theorem 2.1.8 Binomial Formula For any real (or later complex) numbers
a, b and every n ∈ N0 we have
n
n
X n
(a + b) = ai bn−i .
i=0
i
Proof. Again, there are several proofs, and we give two of them.
26 KAPITEL 2. NUMBERS AND FUNCTIONS
(a + b)n+1 = (a + b)n · (a + b)
(IH) Pn n
= i=0 i
ai · bn−i · (a + b)
(DL) Pn n
i+1 n−i
= i=0 i
(a b + ai bn+1−i )
Pn+1 n
i n+1−i Pn n
i n+1−i
= i=1 i−1
ab + i=0 i
ab
(P T ) Pn+1 n+1
i n+1−i
= i=0 i
ab ,
Second Proof: Using combinatorics again. We expand (a + b)n using the distributive law
as a sum
n
X
n
(a + b) = ci ai bn−i .
i=0
(a + b)2 = a2 + 2ab + b2 .
That this fails to go on so nicely with just the binomial coefficients as decimals
is caused by the binomial coefficients becoming larger than 10.
2.2. COMPLEX NUMBERS 27
The set C is closed under addition and multiplication, which means: for all
z = x + yi, w = u + vi, where u, v, x, y ∈ R,
C := R2 = {(x, y) | x, y ∈ R}
and define addition and multiplication by the rules suggested by what we just
saw. For pairs (x, y), (u, v) ∈ C we set
Then one has to check that these operations are commutative and associative and
obey the distributive law (which is tedious and works well) and that – identifying
the real number x with (x, 0) ∈ C leads to the fact that our new operations
extend those for the real numbers, i.e.
This finally shows that our dream came true, and we now set
i := (0, 1)
(x, y) = x · 1 + y · i = x + yi.
This is just the usual vector addition and scalar multiplication, but now we also
can multiply two vectors by the rule invented above.
Every complex number z = x + yi is given by its real part x and its imaginary
part y. We write
Note,
Pn that now, given Q all the properties we wanted in 2.2.1, we may define sums
n
i=0 zi and products i=0 zi as for real numbers and that the binomial formula
will also be true.
wz = w · z and w + z = w + z.
We call this number the absolute value or modulus of z. We see that the absolute
value satisfies a multiplicative property:
√ √ √ √
|zw| = zw · zw = zz · w · w = zz · ww = |z| · |w|, z, w ∈ C.
|z + w| ≤ |z| + |w|, z, w ∈ C.
i.e. multiplying two complex numbers means for the polar coordinates to multiply
the moduli and add the arguments. In particular, the polar coordinates are very
convenient to describe powers of complex numbers:
For instance, every complex number has a square root. This is clear for z = 0,
and for non-zero z = r · (cos(α) + sin(α)i) the possible choices for the square root
are
√
± r · (cos(α/2) + sin(α/2)i).
In particular it is not feasible to establish an order relation on C satisfying all
properties from 1.2.1 because if we had such an ordering a square – and hence
every complex number! – always were ≥ 0.
For example, we would have −1 < 0 due to 0 < 1 and the monotonicity of the
addition and −1 = i2 = (−i)2 > 0, due to the monotonicity of the multiplication,
as i > 0 or −i > 0. Of course this contradiction shows that such an ordering
cannot be established.
Proof. We make use of the geometric properties of the sine- and cosine-function
which hopefully are well-known from highschool.
The proof is based on the following picture, which works well if all angles α, β, α+
β are between 0 and π/2. We hence restrict to this case, the other possibilities
can be dealt with similarly.
The point 0 is the origin and the center of a circle of radius 1 on which the points
P and Q are placed.
...........................
...........
.........
........ Q
......
.........
.... ..... ..............
..
.. .. . ...
...... ..... .....`2......
.. ... ... ....
...... ... ... ...
.. ... . ..
.. .. .......................................0 ...P
.. S . . ..
.
..... β . .... .. ........ ...Q .....
.. ....... .. ...
...... ............... `1 .. ...
.... .......... .. ...
.............. α .. ...
0 ..................................................................................................................
R
The second equation demands v = u−1 , and substituting this into the first gives
after multiplication with u2
u4 + 4u2 − 1 = 0,
which is a quadratic equation in u2 . The formula for solving this tells us that
√
2 −4 ± 16 + 4 √
u = = −2 ± 5,
2
2
but as u has√to be real, u is positive (which still makes sense for real numbers:-)),
hence u2 = 5 − 2. Accordingly,
1 1 √
v2 = 2 = √ = 5+2
u 5−2
√
(as expanding the fraction with 5 + 2 shows).
Our sought for solutions therefore are (as v = u1 again)
√ √
q q
−1
z= · (1 + i ± ( 5−2+( 5 + 2)i)).
2
d
X e
X d+e X
X k
ci x i · bj x j = ( ci bk−i )xk .
i=0 j=0 k=0 i=0
Proof. We write x = x0 +(x−x0 ) and use this together with the binomial formula
2.1.8 to see
2.3. POLYNOMIAL AND RATIONAL FUNCTIONS 35
f (x) = f (x0 + (x − x0 ))
= di=0 ci (x0 + (x − x0 ))i
P
= f (x0 ) + dj=1
Pd i i−j
· (x − x0 )j
P
i=j j x0
Pd Pd i i−j
= f (x0 ) + (x − x0 ) j=1 i=j j x0 · (x − x0 )j−1
where we have singled out in the second last step the summand for j = 0 which
gives f (x
P0 ) and
P then can write the rest as a multiple of (x − x0 ) by a polynomial
g(x) = dj=1 ( di=j ji xi−j
0 ) · (x − x 0 ) j−1
.
This shows that if f (x0 ) = 0, we have f (x) = (x−x0 )·g(x). If, on the other hand,
f is a multiple of (x − x0 ), then of course f (x0 ) is a multiple of (x0 − x0 ) = 0,
hence is zero.
f (x) = (x − x0 ) · g(x),
where g(x) is a polynomial function. Due to 2.3.3, g has degree d and therefore
at most d roots. However, if r ∈ F is a root of f, then
0 = f (r) = (r − x0 ) · g(r),
and this product of two elements of F can only be 0 if at least one of the factors
is zero. Therefore r = x0 or r is a root of g, which together gives at most d + 1
candidates, proving the inductive step.
36 KAPITEL 2. NUMBERS AND FUNCTIONS
(a) Try to find the partial fraction decomposition for the rational function
1
x2 (x−1)2
: z1 = 0, z2 = 1, m1 = m2 = 2.
The numbers ci,j have to satisfy the condition
−(c1,1 + c2,1 )x3 + (2c1,1 − c1,2 + c1,2 − c2,2 )x2 + (−c1,1 + 2c1,2 )x + (1 − c1,2 ) = 0.
This forces c1,2 = 1, hence c1,1 = 2, hence c2,1 = −2 and finally c2,2 = 1.
f (x) 6 7 4
=x+3+ + 2
+ .
g(x) x − 1 (x − 1) (x − 1)3
38 KAPITEL 2. NUMBERS AND FUNCTIONS
Kapitel 3
Sequences
Sequences are a good method to understand the different ways in which one can
approximate complex numbers in a better and better way by other numbers.
3.1 Convergence
Definition 3.1.1 Sequences
A (complex or real) sequence is a map from N (or sometimes N0 ) to the complex
or real numbers.
We denote a sequence often as a = (an )n∈N , where an is the value of the given
sequence a at n. Sometimes we also write a sequence as
a = (a1 , a2 , a3 , . . .) or a = (an )n .
Note, that in contrast to a set, we here have to controll the specific order in which
the values come and that values may occur more often than once – maybe even
infinitely often! For instance, the sequences
and
As every real sequence also is a complex sequence, we will mostly treat these
when talking abut general rules.
39
40 KAPITEL 3. SEQUENCES
This says that for all but finitely many n the values an are located in the disc of
radius around `. We will sometimes abbreviate this by saying that for almost
all n this condition is satisfied.
If such a limit exists, then we write
lim an = `
n→∞
π = 3.1415926535897932384626433832795028842 . . .
and the more decimals you consider the closer you come to π.
But of course there are many more sequences of numbers converging to the same
limit, and this forces us to consider this phenomenon in a more conceptional way.
|` − `0 | = |` − an + an − `| ≤ |` − an | + |an − `0 | ≤ 2 < |` − `0 |,
which of course is a contradiction. Therefore no two different limits for one se-
quence can exist and the limit hence is unique.
Now let ` ∈ C be the limit of (an )n . For every real > 0 we find that for some
N ∈ R all values an , n ≥ N, satisfy |an − `| ≤ /2 because such an inequality
holds for every positive number, and /2 is a positive number. We therefore find
using the triangle inequality again
Now this inequality holds for almost all n and therefore the sequence of differences
an+1 − an is a null sequence.
Note, that the condition that (an+1 − an )n is a null sequence, is not sufficient
for convergence of (an ). We will see a counterexample later in the chapter on
sequences.
Nevertheless, we are now in a good situation to treat our first examples.
42 KAPITEL 3. SEQUENCES
|q n | = (1 + h)n ≥ 1 + nh,
and this will be larger than |`| + 1 as soon as n > |`|/h. Therefore we not
only see that q n cannot converge to `, we even see that |q n | converges to
∞, if |q| > 1.
Our last case now is |q| < 1. Here we note the |1/q| > 1 and therefore
we just saw that |1/q n | →n→∞ ∞, which in turn implies – using our first
example a) in this list – that |q n | →n→∞ 0, proving our last claim.
3.1. CONVERGENCE 43
Then (an )n converges to some complex number if and only if |q| < 1, in
1
which case we have limn→∞ an = 1−q .
It is clear from what we just saw that an+1 − an = q n+1 will only be a
null sequence if |q| < 1, which shows the necessity of this condition for
convergence.
In order to prove the convergence in the case |q| < 1, we recall from 2.1.2
n
b) that an = 1−q
1−q
, and this shows that
1 qn
|an − |=| |.
1−q 1−q
Given any real > 0, we know from example b) that for almost all n we
have |q n | < · |1 − q|, which is just some fixed real number, and therefore
1
for almost all n |an − 1−q | ≤ , proving
n
X 1
q k →n→∞ , if |q| < 1.
k=0
1−q
I again stress that this example – th geometric series – is one of the most
important examples in the whole lecture.
One specific case is q = 0.1. Then the sequence considered starts as
1, 1.1, 1.11, 1.111, 1.1111, 1.11111, . . .
and it is clear that the limit is 1.1. We therefore have shown that
1 10
1.1 = = ,
1 − 0.1 9
which implies
1
0.1 = , hence 0.9 = 1.
9
We already know that from 1.3.6.
an < R.
lim an = `, lim bn = m.
n→∞ n→∞
Before giving the proof we admit that we are a bit careless about an /bn . This
quotient only is defined if bn 6= 0, and we will see in the proof that this is the
case for almost all n, which is what we need for talking about convergence.
Proof. To warm up, we start with an + bn . If any real > 0 is given, we know
that for almost all n
|an − `|, |bn − m| < /2.
Therefore for these n the triangle inequality renders
for almost all n (because /(2R) is real and positive) and therefore by the triangle
inequality
|an bn − `m| = |an bn − `bn + `bn − `m|
≤ |bn | · |an − `| + |`| · |bn − m|
≤ R · /(2R) + R · /(2R)
=
is true for almost all n.
The last assertion concerning the case m 6= 0 uses that for almost all n we have
1 2
|bn − m| < |m|/2, hence |bn − 0| ≥ |m|/2 and | |≤ .
bn |m|
Proof. This is clear from what we said in the last Theorem 3.1.9. As (an ) is
convergent, we find
and so on, i.e. limn→∞ ain = `i for every i ∈ N0 , and then – using the constant
sequence (ci , ci , ci , ci , . . . ) which converges to ci ,
lim ci ain = ci `i .
n→∞
Proof. If a complex sequence (an )n converges to the limit `, then the complex
conjugates (an )n converges to `, because
Therefore the real resp. imaginary parts of an converge to the real resp. imaginary
part of `, as for every z ∈ C
1 1
<(z) = (z + z), =(z) = (z − z).
2 2i
On the other hand, if the real sequences (<(an ))n resp. (=(an ))n converge to x
resp. y , (an )n = (<(an ) + =(an )i)n converges to x + yi.
where
n1 < n2 < n3 < . . .
is a strictly monotonically growing sequence of natural numbers.
For instance, the sequence (1, 3, 5, 7, 9, . . . ) of odd integers is a subsequence of the
sequence a = (1, 2, 3, 4, 5, 6, 7, . . . ) of all natural numbers (both in their natural
order). The sequence (1, 2, 1, 2, 1, 2, . . . ) is no subsequence of a.
It is clear that every subsequence of a convergent sequence converges to the same
limit.
1
|ank − h| < →k→∞ 0.
k
A convergent sequence has exactly one accumulation point. What else can hap-
pen?
48 KAPITEL 3. SEQUENCES
(a) It is easy to construct sequences which have a given finite set as accumulati-
on points and no other accumulation point at all. If S = {s1 , s2 , . . . , sc } ⊆ R
is a finite set, then just take the sequence which starts as
a1 = s 1 , a 2 = s 2 , . . . , a c = s c
1, 0, −1, 2/1, 2/2, 1/2, 0/2, −1/2, −2/2, −2/1, 3/1, 3/2, 3/3, 2/3, 1/3, 0/3,
−1/3, −2/3, −3/3, −3/2, −3/1, 4/1, 4/2, 4/3, 4/4, . . .
Every rational number will be a value of this sequence at least once, even
infinitely often. Therefore every rational number is an accumulation point.
Even worse (or even more interesting, depending on your taste): As for every
real number x and every > 0 there are infinitely many rational numbers
in (x − , x + ) by the argument given in 1.3.5, x is an accumulation point.
The set of accumulation points is the whole real line!
Note, that this argument also shows that Q cannot be the set of all accu-
mulation points of a sequence.
We already saw that every convergent sequence is bounded. The converse is false,
as you can see from the sequence ((−1)n )n∈N . But nevertheless the following
theorem is very important.
3.2. SUBSEQUENCES AND ACCUMULATION POINTS 49
Proof. If we know the theorem for real sequences, then it follows for complex
sequences as well, because for a bounded complex sequence (an )n the real- and
imaginary parts are also bounded. The bounded real sequence of real parts has a
convergent subsequence (<(ank ))k and then the bounded real sequence (=(ank ))k
again has a convergent subsequence. The corresponding subsequence of (an )n
therefore converges due to 3.1.11.
We may therefore assume that (an )n is a bounded real sequence, the values being
contained in the interval (−R, R). After adding R and dividing by 2R we may
even assume 0 ≤ an ≤ 1 for all n.
We now have to “find” an accumulation point, which we “construct” as a decimal
number – this is an existence proof! We cannot really single out an accumulation
point explicitely.
We exclude the case that infinitely many members of the sequence are 1 – in this
case, 1 were an accumulation point and we were done with.
We now start by choosing the maximal integer b1 such that for infinitely many
n
b1
≤ an
10
This b1 exists because the set
{z ∈ Z | z ≤ 10an for infinitely many n}
is not empty (it contains every integer z < 0 ) and bounded (no such z is larger
than 9 ).
More generally we define for k ∈ N the integer bk to be maximal with the
property that for infinitely many n
bk ≤ an · 10k .
Again, it is clear that these bk exist by the same reason. As moreover 10bk ≤
an · 10k+1 infinitely often but 10(bk + 1) > an · 10k+1 for almost all n (due to the
choice of bk ), we see that
10bk ≤ bk+1 < 10bk + 10.
Therefore there exists an integer dk+1 ∈ {0, 1, . . . , 9} such that
bk+1 = 10bk + dk+1 .
This defines a decimal number x = 0.d1 d2 d3 . . . = lim∞
k=1 dk 10
−k
, and ` = b0 +x =
bk
limk→∞ 10k .
50 KAPITEL 3. SEQUENCES
bk bk + 1
k
≤`≤ ,
10 10k
and as also an is between these two bounds infinitely often, we get
Proof. It is clear that if (an )n converges there is exactly one accumulation point,
namely its limit.
To show the converse we give a proof by contradiction.
Assume that (an )n has eactly one accumulation point ` which is not the limit
of (an )n . Then there exists an > 0 such that for infinitely many n
|an − `| > .
Proof. We only deal with the case that an ≤ an+1 for all n, the other case being
very similar.
As the sequence is bounded, there exists an accumulation point. This accumula-
tion point is unique, as one can see by contradiction. To that end, assume that
h < h0 are two accumulation points. Then there exists an N ∈ N such that
which implies
aN ≥ h + (h0 − h)/2.
As all later members of the sequence are larger than aN , we have
and therefore h cannot be an accumulation point. 3.2.6 then gives the desired
convergence.
We now are in good shape to treat some examples which will turn out to be useful
later.
Qn := Fn /Fn−1 , n ∈ N.
It is desirable to prove convergence of this, because then by 3.1.9 we see that the
limit ` satisfies
` = 1 + 1/`,
i.e. √
2 1± 5
` − ` − 1 = 0, i.e. ` = ,
2
√
1+ 5
and as Qn always is positive, we have ` = 2
.
Let us write down the first members of the sequence (Qn )n .
Q1 = 1, Q2 = 2, Q3 = 3/2, Q4 = 5/3, Q5 = 8/5, Q6 = 13/8, . . .
We see that
Q1 < Q3 < Q5 < Q6 < Q4 < Q2 .
Observe that
2Qn + 1
Qn+2 = 1 + 1/Qn+1 = .
Qn + 1
As now for all x > y > −1 we find 2x+1 x+1
> 2y+1
y+1
, it follows from Q1 < Q3
that Q3 < Q5 < Q7 < Q9 . . . , i.e. the subsequence (Q2n−1 )n∈N is monotonically
growing. It of course is bounded, because Qn ≥ 1 implies Qn+2 = 1 + 1/(1 +
Q−1n ) ≤ 2. Similarly Q2n is decreasing monotonically because Q2 > Q4 > Q6 >
. . . , and is bounded from below by 1. Therefore both subsequences converge by
the monotonicity criterion 3.2.8 and using 3.1.9 for the subsequences we see that
the limit `odd = limn→∞ Q2n−1 satisfies
2`odd + 1
`odd = ,
`odd + 1
2x+1
and the same equation holds for `even = limn→∞ Q2n . The equation x = x+1
,
however, gives after multiplication by 1 + x
x + x2 = 2x + 1, i.e. x2 − x − 1 = 0,
which in the end shows what we wanted.
This shows that no two different accumulation points can exist. A similar argu-
ment will later be used in the context of “absolute convergence” of series.
We now define the function exp : R → R by
n
X xk
exp(x) = lim .
n→∞
k=0
k!
This is called the exponential function and will play an important role again and
again.
In particular one sets e := exp(1) and can calculate good approximations to
this number – being called Euler’s number – as the sequence converges rapidly
towards its limit:
n
X 1
e = lim = 2.718281828459045 . . .
n→∞
k=0
k!
Unfortunately we can not yet justify the wording, but will do so later.
Kapitel 4
Continuity
55
56 KAPITEL 4. CONTINUITY
We then write
lim f (s) = `,
s→z0
where we tacitely imply that all s under consideration lie in S, so that f (s) is
defined.
Of course there are examples, where the said limit does not exist at all. For instan-
ce limx→0 sin(1/x) does not exist, as, e.g., the limit of the sequence sin(1/xn )n
2
with xn = πn does not exist.
If z0 ∈ S is an accumulation point of S then f is continuous at z0 if and only
if lims→z0 f (s) = f (z0 ).
This lemma gives a precise meaning to the vague statement that f is continuous
at z0 if f (z) is close to f (z0 ) as soon as z ∈ D is close enough to z0 .
a) The function
| · | : C → R, z 7→ |z|,
is continuous. Because: ||z| − |w|| ≤ |z − w| by 1.2.4. This means: For fixed
z0 and any sequence (sn )n converging to z0 we find that (|z0 | − |sn |)n is a
null sequence, whence sn converges to z0 .
b) The function
0 if x ∈ Q,
f : R → R, f (x) =
1 if x 6∈ Q,
is nowhere continuous. Because for every x0 ∈ R and every δ > 0, there
are rational and irrational numbers in the interval (x0 − δ, x0 + δ) due to
1.3.5, and therefore in every such interval the values 0 and 1 are taken by
f infinitely often.
Similarly, for this f, the function g(x) = f (x) · x2 is continuous only at
x0 = 0, and nowhere else.
c) From 3.1.10 we remember that – without having had the terminology then
– we verified that polynomial functions always are continuous everywhere.
d) From 3.1.9 we see that sums and products of continuous functions defined
on the same domain are continuous again. The same for quotients, of the
divisor does not take the vaue 0 anywhere.
Proof. We prove the existence of the maximum. The minimum goes similarly. As
C is bounded, there exists an upper bound b0 of C. As C is non-empty, there
is an element c0 ∈ C. We have c0 ≤ b0 and successively define two sequences:
(bn )n will be a decreasing sequence of upper bounds of C and (cn )n an increasing
sequence of elements of C such that |bn − cn | is a null sequence.
If b0 ≥ b1 ≥ . . . ≥ bn and c0 ≤ c1 ≤ . . . ≤ cn are already defined, we look at
x = an +b
2
n
. If x is an upper bound for C, we set cn+1 = cn , bn+1 = x. If it is
not an upper bound for C, there exists a cn+1 ∈ C with x < cn+1 and we set
bn+1 = bn . In both cases |bn+1 − cn+1 | ≤ 21 |bn − cn |, which makes the difference
a null sequence.
As both sequences are monotonic and bounded, they converge by the monotoni-
city criterion 3.2.8, and as their difference is a null sequence, they converge to the
same element M. But as C is closed and cn ∈ C, we have M ∈ C. On the other
hand, as M is a limit of upper bounds, it also is an upper bound. Therefore, M
is the maximum of C.
The miracle now is that the range of a continuous function on a compact domain
is compact.
Boundedness: Assume that f (C) is not bounded. Then for each n ∈ N there
exists an element cn ∈ C with |f (cn )| ≥ n. As C is bounded, (cn )n has a
convergent subsequence by Bolzano-Weierstraß, 3.2.5. Without loss of generality
we may assume (cn )n to be convergent. As C is closed, the limit ` = limn cn
also belongs to C, but then limn f (cn ) = f (`), which contradicts |f (cn )| ≥ n >
|f (`)| + 1, which holds for large n. This shows boundedness of f (C).
Closedness: If (an )n is a convergent sequence in f (C), we choose cn ∈ C such
that f (cn ) = an for every n. Again, (cn )n contains a convergent subsequence
(cnk )k , and then
Another basic property of continuous functions is special for the real case. To
formulate it smoothly we say that a real number x lies between two other real
numbers y, z if y ≤ x ≤ z or z ≤ x ≤ y, i.e. x ∈ [min(y, z), max(y, z)].
Define
S := {x ∈ [a, b] | f (x) ≤ v}.
Then S is closed, because if (sn )n is a convergent sequence in S with limit `,
continuity implies f (`) = limn f (sn ) ≤ v, i.e. ` ∈ S. As S ⊂ [a, b], S is bounded
and hence it is compact and therefore has a maximum x0 by 4.2.2. As x0 ∈ S,
we have f (x0 ) ≤ v.
As v < f (b), there are sequences (tn )n in (x0 , b] converging to x0 . But tn > x0
implies f (tn ) > v, and therefore – by continuity again – f (x0 ) = limn f (tn ) ≥ v.
This shows f (x0 ) = v.
Series
for every n, and it sometimes is very helpful to write sequences in this way, with
given numbers di . This leads to the definition of series.
the n -th partial sum of (an )n . The sequence of partial sums (sn )n is called the
(infinite) series defined by (an )n , and if it converges to a limit `, we write
∞
X
`= ai .
i=0
63
64 KAPITEL 5. SERIES
converges for every x ∈ R (cf. 3.2.11). We will see in 5.1.5 combined with
5.1.6 that it even converges for every x ∈ C.
P∞ i
(b) The geometric series n=0 q converges if an only if |q| < 1, as we saw in
3.1.6. For such q we now may write
∞
X 1
qi = .
i=0
1−q
P∞ 1
(c) The harmonic series n=1 n does not converge, although the summands
are a null sequence.
Pn 1 This is a consequence of the fact that for the partial
sums sn = k=1 k we see that
2n
X 1 1 1
s2n − s2n−1 = ≥ 2n−1 n = ,
k 2 2
k=2n−1 +1
Proof. Let sn = nk=0 ak be the n -th partial sum of (ak )k and tn = nk=0 |ak |
P P
the n -the partial sum of (|ak |)k .
As (tn )n converges, the (real) sequence (tn ) is bounded by some R > 0. By the
triangle inequality,
n
X n
X
|sn | = | an | ≤ |ak | = tn ≤ R
k=0 k=
|` − sn |, |`0 − sm | ≤ R/3,
resulting in
|sn − sm | ≥ R/3
infinitely often, which we just disproved.
As (sn )n is bounded and has exactly one accumulation point, it is convergent by
3.2.6.
66 KAPITEL 5. SERIES
∞
X X∞ X∞
cn = ( an ) · ( bn ).
n=0 n=0 n=0
P∞ P∞
Proof. Let R be a common upper bound for n=0 |an | and n=0 |bn |.
Then for every k,
k
" k
# " k
#
X X X X
|cn | ≤ |am | · |bn | ≤ |am | · |bn | ≤ R2
n=0 m,n:m+n≤k m=0 n=0
We know from 3.1.9 that the product on the right hand side of the inequality in
the statement is the limit of the products of the partial sums:
X∞ X∞ Xk Xk X
( an ) · ( bn ) = lim ( an ) · ( bn ) = lim am b n .
k→∞ k→∞
n=0 n=0 n=0 n=0 0≤m,n≤k
We get
2k
X k
X X
| cn − am b n | = | am bn |,
n=0 n,m=0 (m,n)∈Tk
where
Tk = {(m, n) ∈ N20 | m + n ≤ 2k and m > k or n > k}.
5.1. INFINITE SUMS 67
P∞ P∞
For given > 0 and large enough k, m=k |am | ≤ /(2R) and n=k |bn | ≤
/(2R) Therefore, using the triangle inequality, we have
P P
| (m,n)∈Tk am bn | ≤ (m,n)∈Tk |am | · |bn |
P∞ P∞
≤ 2k
P P2k
m=0 |am | n=0 |bn | + m=0 |am | n=0 |bn |
≤ 2 · R · /(2R) = ,
showing in combination with the last inequality that the two sequences have the
same limit.
converges absolutely.
Using the Cauchy-Convolution formula and the Binomial Formula 2.1.8 this im-
plies for all z, w ∈ C
exp(z) · exp(w) = ∞
P zm
P∞ wn
m=0m! ·
P∞ Pk n=0 n!
z m wk−m
= k=0 m=0 m! (k−m)!
P∞ Pk k
m k−m
= k=0 m=0 m z w /(k!)
P∞ (z+w)k
= k=0 k!
= exp(z + w).
This identity is called the functional equation of the exponential function. We will
use it again and again.
First of all it shows that for natural numbers d
d
X
exp(d) = exp( 1) = exp(1)d = ed ,
i=1
We also see that on real numbers the exponential function is monotonically in-
creasing: for all real numbers x < y,
For complex arguments this is meaningless, and we will even see that on C every
value of exp is non-zero and taken infinitely often.
as ak+2 ≤ ak+1 .
For odd k we find
sk+2 = sk + ak+1 − ak+2 ≥ sk
for the same reason.
As a0 = s0 ≥ 0 and s1 = a0 −a1 ≥ 0, we see for odd k, due to this monotonicity,
sk ≥ 0. As then sk+1 = sk + ak+1 this is also true for all even k. Similarly, for
all even k we see sk ≤ a1 , and this then also holds for sk+1 = sk − ak+1 .
Therefore, all partial sums sk satisfy
0 ≤ s k ≤ a0 ,
converge. As their difference is a Null sequence (only here we need that an con-
verges to 0), they have the same limit. As in 3.2.2 we now see that this limit is
the only accumulation point ` of (sk )k and therefore sk converges.
Due to the monotonicities of the odd and even subsequences of partial sums, `
lies between sk and sk+1 , showing
|` − sk | ≤ |sk+1 − sk | = ak+1
as desired.
Proof. Looking at the sequences of partial sums, this is clear from 3.1.9.
|an+1 |
Proof. Choose N ∈ N such that |an |
≤ q for all n ≥ N.
Then by mathematical induction for all k ∈ N0 we have
|aN +k | ≤ q k |aN |
which shows absolute convergence by 5.2.1, using the convergence of the geometric
series.
|an | ≤ q n
for almost all n, which shows absolute convergence by 5.2.1, using the conver-
gence of the geometric series.
The most important application of these theorems will be the topic of the next
section.
where cn ∈ C are the coefficients and x is the variable. The set of all x for which
this series converges is called its domain of convergence.
72 KAPITEL 5. SERIES
Proof. As f (z) converges, the sequence (|cn z n |)n is a null sequence, and in par-
ticular it is bounded from above by some R > 0.
As |w| < |z|, we have q := |w/z| < 1. Then
pn
p √
n
|cn wn | = n |cn z n |q n ≤ R · q →n→∞ q < 1
and hence
p
n 1+q
|cn wn | ≤ < 1 for almost all n
2
shows absolute convergence by the Root Test 5.2.5.
Proof. We have to show that the function g(h) := f (z0 + h) can be written as a
power series in h (if |h| is small enough).
Let h ∈ C have absolute value |h| < ρ−|z0 |. Then f (z0 +h) converges absolutely
and we have
∞ ∞ n
!
X X X n
f (z0 + h) = cn (z0 + h)n = cn · hk z0n−k .
n=0 n=0 k=0
k
As this converges absolutely, we may rearrange the summands and put together
summands involving the same power of h, i.e.
∞ ∞ ! ∞
X X n n−k k
X
f (z0 + h) = cn z0 ·h = c̃k hk ,
k=0 n=k
k k=0
P∞ n
n−k
where c̃k = n=k cn k z0 .
We gain from the proof that c̃k converges, but this also could be established
independently from the absolute convergence of c̃0 = f (z0 ).
As N n
P
n=0 cn z is a polynomial in z we know that there is some positive δ which
may be chosen smaller than s such that
N
X
for all z with |z| < δ : | cn z n − f (0)| < /2.
n=0
PN n
This comes from the fact that f (0) = n=0 cn 0 .
Putting together these two estimates (and using the triangle inequality) gives the
desired
|f (z) − c0 | ≤ 2/2 = if |z| < δ.
It is clear from the definition that S(−z) = −S(z) and C(−z) = C(z). This
also follows from the fact that in C only even powers of z are involved, while in
S only odd powers are involved.
The definition of S and C and the identity exp(z) = exp(−z)−1 leads to
For real x, S(x) and C(x) are real, because the coefficients in the power series
are real. For two real numbers x, u we use the Functional Equation of exp , 5.1.8,
in the following calculation:
This also holds for complex x and u and is the same addition formula as for the
cosine and sine function from 2.2.6.
We will see later in 6.1.5 that C and S actually are the cosine and sine function.
The next lemma will be one step further in this direction.
Proof. For every x ∈ [0, 2] the sequence (x2n /(2n)!)∞ n=1 is positive and monoto-
nically decreasing. Therefore, using the last assertion of Leibniz’ Criterion 5.1.10,
we find that
|C(x) − (1 − x2 /2)| ≤ x4 /24,
showing that C(2) < 0, while C(0) = 1 > 0. By the Intermediate Value Theorem
4.2.7 we know that the continuous function C has at least one zero in [0, 2].
With a similar argument we find that for all x ∈ [0, 2],
S(x) ∈ [x − x3 /6, x]
76 KAPITEL 5. SERIES
exp(ηki) = ik .
A special case is exp(℘i) = −1. Later on we will see that ℘ = π and then this
last equation is Euler’s equation.
As consequently exp(2℘i) = 1, we – again by the functional equation – derive
that the exponential function is periodic:
as well as
S(z + ℘) = −S(z), C(z + ℘) = −C(z).
5.4. THE EXPONENTIAL FUNCTION AND SOME RELATIVES 77
and in particular | exp(x + yi)| = exp(x). Remember that for real numbers x > 0
we have exp(x) > 1 (just as every summand in the power series is positive) and
1
for real numbers x < 0 we have exp(x) = exp(−x) < 1.
This implies that for solving exp(z) = 1 you need <(z) = 0 and C(=(z)) = 1.
As S(y) > 0 for y ∈ (0, 2) by 5.4.2 and for y ∈ (℘ − 2, ℘) by the other results,
℘ is the smallest positive zero of S. As the zeros of S have period ℘, we must
have =(z) ∈ Z · ℘, showing z = k℘i. For odd k, exp(z) = −1. Hence we have
exp(z) = 1 ⇔ z ∈ Z · 2℘i.
As already said, all that points pretty much in the direction that C = cos, S = sin
and ℘ = π, but we do not know that yet.
The missing step is that we have to show, that the arc length covered by the
curve t 7→ (C(t), S(t)), 0 ≤ t ≤ x, is x. This curve has values on the unit circle,
because S 2 + C 2 = 1. As already indicated, this gap will be closed in 6.1.5.
Proof. The last assertion is clear by what we just saw as exp(z) = exp(w) if and
only if exp(z − w) = 1.
As exp(z) · exp(−z) = 1, 0 does not belong to the range of exp .
We now have to show that every non-zero complex number w does belong to the
range. Write
w = r · (x + yi), r > 0, x, y ∈ R, x2 + y 2 = 1.
For every real t > 0, exp(t) > 1 by the power series defining exp . Therefore,
using 3.1.6(b), the sequence of numbers exp(nt) = exp(t)n , n ∈ N is unbounded.
By the intermediate value theorem, 4.2.7, every real number r ≥ 1 can therefore
be written as r = exp(t) for some t ∈ [1, ∞). If 0 < r < 1, there is a t ≥ 1 with
1
r
= exp(t), i.e. r = exp(−t).
Note that exp is strictly monotonically increasing on R by the functional equa-
tion.
Now we have to find a real u with exp(ui) = (x + yi).
78 KAPITEL 5. SERIES
We know that C takes every x ∈ [0, 1] as a value between 0 and ℘/2. Due
to C(℘ − u) = −C(u) we see that C takes every x ∈ [−1, 1] as a value on
[0, ℘] . Having found u with C(u) = x, we know that S(u) = ±y, because
S(u)2 = 1 − C(u)2 = 1 − x2 = y 2 .
As (C(−u), S(−u)) = (C(u), −S(u)) we finally find a suitable u ∈ [−℘, ℘] with
(C(u), S(u)) = (x, y), if x2 + y 2 = 1.
We can write down the power series giving these functions, inserting those for
the exponential function in the defining equation:
∞ ∞
X z 2n X z 2n+1
cosh(z) = , sinh(z) = .
n=0
(2n)! n=0
(2n + 1)!
cosh(z)2 − sinh(z)2 = 1,
80 KAPITEL 5. SERIES
which is responsible for the fact that we can parametrize the real hyperbola
{(x, y) ∈ R | xy = 1}
as
{(cosh(t) + sinh(t), cosh(t) − sinh(t)) | t ∈ R}.
This justifies the names hyperbolic cosine and hyperbolic sine.
In 7.4.4 we will sketch why the hyperbolic cosine is the function describing a
catenary.
One can even define a hyperbolic tangens and the like as in the trigonometric
case.
Proof. Taking differences, we have to show that the only way to represent the
zero function on (−r, r) by a power series is with all coefficients being zero.
Assume now that
∞
X
0= cn xn for all x ∈ (−r, r).
n=0
Looking at x = 0 we find c0 = 0.
Now assume further, that not all coefficients are zero, and let m be the smallest
index with cm 6= 0. We want to derive a contradiction.
Write the right hand side as
∞
X
xm · cm+n xn .
n=0
xm 6= 0 for x 6= 0, we have
∞
X
cm+n xn = 0 for all x ∈ (−r, r), x 6= 0.
n=0
But as this power series is continuous, it has to be zero on the whole interval
(−r, r), giving constant term 0 again: cm = 0.
This contradicts our assumption and shows that all coefficients are zero.
We already knew this for the special case of polynomials, which are finite power
series, and where we saw the uniqueness of coefficients in 2.3.5: A non-zero poly-
nomial cannot vanish on a whole interval.
In order to calculate at least the first 4 coefficients, you can truncate this after
n = 4 and neglect all higher powers of x. Setting g(x) = c1 x + c2 x2 + c3 x3 + c4 x4
this leads to
1 4
+ ( 24 c1 + 21 c2 c21 + c3 c1 + 21 c22 + c4 )x4 + higher terms.
Differentiability
f (x) − f (x0 )
lim
x→x0 x − x0
83
84 KAPITEL 6. DIFFERENTIABILITY
exists. If this is the case, the limit is called the derivative of f at x0 , denoted by
f 0 (x0 ).
The quotient studied in the limit is called the difference quotient, its limit – if it
exists – the differential quotient of f at x0 .
Often one calculates the derivative as
f (x0 + h) − f (x0 )
lim , 0 6= h ∈ D − x0 .
h→0 h
This notation just replaces x by x0 + h, and of course one is interested in values
of h such that 0 < |h| is small enough so that x0 + h ∈ D holds.
If the derivative exists on all of D, we say that f is differentiable on D. If then
the derivative again is differentiable, its derivative is called the second derivative
f 00 of f.
Recursively, one defines the (n + 1) -st derivative of f as the derivative of the
n -th derivative, if this exists. The n -th derivative is denoted as f (n) 1
Sometimes, in particular if f also depends on some parameters, we write
df
(x0 ) = f 0 (x0 ).
dx
This will become the prefered notation for functions in more than one variable.
If in physical applications the variable is time, then it often is denoted by t and
the derivative of f with respect to time is f˙ (yes, there is a dot over the f ).
This idea of best approximation will be taken up again in the section on Tay-
lor polynomials. We will also use it in the proof of the chain rule and in some
examples.
...
1
and not as f 000 :-)
6.1. THE DERIVATIVE 85
We note, however, that there could be other criteria for the quality of an approxi-
mation, and that this pretty much depends of what you need the approximation
for.
hence this is a power series which is continuous and evaluates to its constant
term c1 at x = 0 :
f 0 (0) = c1 .
86 KAPITEL 6. DIFFERENTIABILITY
(c) We apply this last insight to f (x) = exp(x) and see exp0 (0) = 1. Now we
look at the functional equation of the exponential function and calculate
for h 6= 0
exp(x0 +h)−exp(x0 )
h
= exp(x0 )(exp(h)−1)
h
= exp(x0 ) exp(h)−exp(0)
h−0
→h→0 exp(x0 ) · exp0 (0)
= exp(x0 ).
The exponential function is its own derivative!
For the functions S and C from 5.4.1 we see from (b) that S 0 (0) = 1 and
C 0 (0) = 0 and again, using the addition formula from the same number
that
S 0 (x) = C(x), C 0 (x) = −S(x).
(a) In physics (or other sciences) the movement of a particle in space during
some time interval [0, T ] is given by a map
As we will only talk about integrals in the following chapter, we just point
out that if the speed does not change, i.e. kḞ (t)k = s for fixed s and every
t ∈ [0, T ] you get
L = T · s.
Similar rules hold for a movement in the plane, which of course can be
viewed as a part of three dimensional space.
(b) Now look at the functions S and C from 5.4.1 again. We just learned that
S 0 (t) = C(t), C 0 (t) = −S(t) and already know that S(t)2 + C(t)2 = 1
for every real t. As we look at a “particle” moving along the unit circle
described by
F : [0, T ] → R2 , F (t) = (C(t), S(t)),
this particle starts at (1, 0) and moves up to some point on the unit circle,
which we now can describe in two ways:
for suitable α ∈ R. As the measurement for angles is the arc length (up
to multiples of 2π ) we derive from this that – as the speed of the particle
constantly is 1 –
T = α + 2πk, k ∈ Z suitable.
∞ ∞
X t2n+1 X t2n
sin(t) = (−1)n and cos(t) = (−1)n , t ∈ R.
n=0
(2n + 1)! n=0
(2n)!
We could have derived this from the addition formula, given that we believe
that sin0 (0) = 1, but now we have proved it, although much more indirect.
The motivation for finding the power series describing sine and cosine is
outlined in the section on Taylor series, 6.4.10.
88 KAPITEL 6. DIFFERENTIABILITY
(a) We have
Observe that
where
t(h) = f 0 (x0 )g 0 (x0 )h2 +(f (x0 )+f 0 (x0 )h)·s(h)+(g(x0 )+g 0 (x0 )h+s(h))·r(h).
lim t(h)/h = 0,
h→0
(c) We now first assume that g is differentiable and non-zero at x0 and that
also g1 is differentiable at x0 . Then the product formula applied to 1 =
1
g(x) · g(x) renders the equality (as 1 is the constant function and clearly
has derivative zero)
0
0 0 1 1
0 = 1 = g (x0 ) · + g(x0 ) (x0 ),
g(x0 ) g
giving 0
1 −g 0 (x0 )
(x0 ) = .
g g(x0 )2
Using the product formula again, we derive from this formula the one clai-
med for the derivative of f /g.
What remains is to prove the differentiability of 1/g at x0 . We leave this
as an exercise.
where t(h) = g 0 (f (x0 ))r(h) + s(f 0 (x0 )h + r(h)). As f 0 (x0 )h + r(h) goes to zero
as h goes to zero, we have
s(f 0 (x0 )h + r(h)) s(f 0 (x0 )h + r(h)) f 0 (x0 )h + r(h)
lim = lim · = 0,
h→0 h h→0 f 0 (x0 )h + r(h) h
because the first factor goes to 0, while the second factor goes to f 0 (x0 ).
Here we may tacitely only look at the case f 0 (x0 )h + r(h) 6= 0, because s(0) = 0
anyway. This shows together with r(h)/h →h→0 0 , that also
lim t(h)/h = 0,
h→0
Proof.
We use the chain rule 6.2.3 in order to explain the formula giving (f −1 )0 (y0 ).
We do not prove differentiability formally, which could be done similarly to the
differentiability of a composition.
From y = f (f −1 (y)) for all y ∈ E we find from differentiating both sides
1 = f 0 (f −1 (y0 )) · (f −1 )0 (y0 ),
which implies the claim because f −1 (y0 ) = x0 .
0 cos(x)2 + sin(x)2
tan (x) = 2
= 1 + tan(x)2 .
cos(x)
This gives then
1
arctan0 (y) = .
1 + y2
f (x) ≤ f (x0 ).
Proof. We treat the case of a local point of minimum, the other case being similar.
By definition of local points of minimum, in a (sufficiently small) subinterval
(x0 − r, x0 + r) ⊆ (a, b) the numerator of the difference quotient
f (x) − f (x0 )
Q(x) = , x 6= x0 ,
x − x0
always is non-negative, the denominator changes sign.
This shows Q(x) ≤ 0 for x < x0 and Q(x) ≥ 0 for x > x0 .
Therefore f 0 (x0 ) = limx→x0 Q(x) = 0.
This lemma implies a pre-version of the mean value theorem.
Proof. As the interval [a, b] is closed and bounded, we can apply 4.2.5 and deduce
that f does attain a maximal and a minimal value on [a, b]. If f (a) = f (b) is
both maximal and minimal, the function is constant and therefore the derivative
vanishes at every point.
In the other case, there must be a point of extremum x0 ∈ (a, b). As every point
of extremum also is a local point of extremum, by 6.3.2 we have f 0 (x0 ) = 0 at
this point.
f (b) − f (a)
f 0 (x0 ) = .
b−a
Proof. We modify f slightly to get a function for which we can apply Rolle’s
Theorem 6.3.3.
To that end, let
F (x) = f (x) − λ(x − a).
We choose λ ∈ R such that
As F clearly still is continuous on [a, b] and differentiable on (a, b), we find some
x0 ∈ (a, b) with F 0 (x0 ) = 0, which, by the rules for taking derivatives, leads to
0 = F 0 (x0 ) = f 0 (x0 ) − λ,
which is the desired equality.
With the same argument, f 0 (x) ≤ 0 (resp. < 0 ) on I ◦ will lead to f being
monotonically decreasing (resp. strictly monotonically decreasing).
Proof. We only consider the case f 00 (x0 ) < 0, the other case being quite similar.
As due to the continuity f 00 (x) is negative in a small interval (x0 − r, x0 + r)
around x0 , we may apply 6.3.5 and see that f 0 is strictly decreasing on this
intervall.
In particular, f 0 (x) > 0 for x ∈ (x0 − r, x0 ) and f 0 (x) < 0 for x ∈ (x0 , x0 + r).
Therefore, again by 6.3.5, f is strictly monotonically increasing on (x0 − r, x0 ]
and decreasing on [x0 , x0 + r), showing the (local) maximality of f (x0 ).
6.3. THE MEAN VALUE THEOREM AND EXTREMAL VALUES 95
(a) We already know from 6.2.2 that a polynomial function of degree d satisfies
f (d+1) = 0.
Let now I ⊆ R be an interval and f : I → R a function which is d + 1
times differentiable and satisfies f (d+1) = 0.
We know from 6.2.2 that every polynomial function is the derivative of a
polynomial function hence from 6.3.5(a) that every function on the interval
I which has a polynomial as its derivative, is itself a polynomial.
Repeating this process several times, starting with f (d+1) = 0, we see that
f (d) is a polynomial and then f (d−1) is a polynomial and so on, showing
that f itself is a polynomial function of degree at most d.
(G − G̃)0 = G0 − G̃0 = g − g = 0.
There is still another variant of the mean value theorem we should look at.
for this x.
Proof. Again, we invent a new function F for which we can apply Rolle’s Theorem
6.3.3. This now is
We find F (a) = f (b)g(a) − g(b)f (a) = F (b) and therefore Rolle’s Theorem gives
us the x we wanted.
96 KAPITEL 6. DIFFERENTIABILITY
If here g 0 is nowhere zero, again Rolle’s Theorem implies that g(b) 6= g(a) and
therefore we can divide the whole equation by (g(b) − g(a)) · g 0 (x), giving the
second equation.
There are two interesting consequences of this variant. One of them will be consi-
dered in treating the remainder term of Taylor’s polynomials in the next section.
The other one is:
Proof. We know from 6.3.8 that for every x ∈ (a, b) there exists some x̃ ∈ (a, x)
such that
f (x) − f (a) f 0 (x̃)
= 0 .
g(x) − g(a) g (x̃)
If x goes to a, the corresponding x̃ also goes to a, because it sits between a
and x.
Therefore the right hand side converges to a limit, as x → a, but the left hand
side has the same values, forcing it also to converge to the same limit. As f (a) =
g(a) = 0 this just is what we wanted.
But then
x2 2x
lim x
= lim x = 0,
x→∞ e x→∞ e
ex ex
lim x = lim ex x = 0.
x→∞ ee x→∞ e · e
x
And then of course there are functions which grow much faster than ee . . .
x
lim (1 + )t = ex .
t→∞ t
This is implied by
x ln(1 + xt )
1
1+ xt
· −x
t2
lim t ln(1 + ) = lim = lim −1 = x,
t→∞ t t→∞ 1/t t→∞
t2
where we did use L Hôpital’s rule in between, taking the derivative with
respect to t of course ( x being fixed).
Of course, this identity can now again be used to calculate
1 n
e = lim (1 + ) .
n→∞ n
98 KAPITEL 6. DIFFERENTIABILITY
Proof. We make use of the calculations in 5.3.5 and remind us of the fact that
for fixed x0 ∈ (−r, r) and small enough h we have
∞
X
f (x0 + h) = c̃k hk ,
k=0
P∞ n
n−k
where c̃k = n=k cn k x0 . This power series in the
P variablen−1h now is diffe-
rentiable by 6.1.4(b) at h = 0 with derivative c̃1 = ∞
n=1 cn nx0 . Replacing n
with k = n − 1 gives the formula from the statement.
f 0 (x) = x · f (x)
which should hold now for all real x. If you do not have an idea how to deal with
this, then just try to write f as a power series around x0 = 0 :
∞
X
f (x) = ck x k ,
k=0
which leads to
∞
X ∞
X
0 k
f (x) = (k + 1)ck+1 x = xf (x) = ck−1 xk ,
k=0 k=1
i.e.
c1 + 2c2 x + 3c3 x2 + 4c4 x3 + · · · = c0 x + c1 x2 + c2 x3 + c3 x4 + . . .
100 KAPITEL 6. DIFFERENTIABILITY
Now, according to 5.5.1, for every k the coefficients for xk have to coincide,
because both power series give the same function. Therefore,
And indeed, plugging in this function into the equation we wanted to solve, we
see that is does satisfy
d
exp(x2 /2) = exp(x2 /2) · x
dx
by the chain rule.
We could have guessed this solution, but sometimes it is good not to depend on
ones capacities to guess something.
How should we have guessed this? If you want to solve the differential equation
for some fixed function g (a potential or something), and if this “potential” has
an antiderivative G, i.e. G0 (x) = g(x) for all x, then at least one type of solutions
is of the form
f (x) = c0 · exp(G(x))
again using the chain rule. And of course G(x) = x2 /2 is an antiderivative of
g(x) = x.
Whether these are all solutions for this particular equation or whether there
are other solutions at all is a totally different question which we will address in
7.4.5(b).
f (k) (x0 ) = 0, 0 ≤ k ≤ n.
Then for every x ∈ (a, b) there exists a z between x0 and x such that
This remainder has the property that it is differentiable n times (as f and pn
(k)
are) and Rn (x0 ) = 0 for 0 ≤ k ≤ n. Therefore, if f is even differentiable (n+1)
times, we find for every x ∈ I a z between x0 and x such that
f (n+1) (z)
Rn (x) = · (x − x0 )n+1 .
(n + 1)!
k 0 1 2 3 4 5
a
k
1 12 − 18 1
16
5
− 128 7
256
Comparing that to the values we started of with in 5.5.2 should persuade you
that Taylor series are good to have. They facilitate the work we had there and
make it much more elegant. Lagrange’s form of the remainder term tells us that
n 1
√ 1
1
X
1+x− 2 k
x = 2 z 2 −n−1 xn+1
k=0
k n+1
for some z between 1 and 1 + x. One can use this estimate to show that for
x ∈ (−1, 1) we get
∞ 1
√ X
1+x= 2 xk .
k=0
k
This results in
cos(2n) (0) = (−1)n = sin(2n+1) (0), n ∈ N0 ,
while the other derivatives are zero. The corresponding Taylor series therefore
are the power series representing cos and sin from 6.1.5(b).
104 KAPITEL 6. DIFFERENTIABILITY
f (k) (0) = 0, k ∈ N0 .
Here, arctan is the “arcus tangens”, inverting the tangens restricted to the in-
terval (−π/2, π/2).
Kapitel 7
Integration
105
106 KAPITEL 7. INTEGRATION
It is quite natural to demand that the second wish extends to infinite unions in
the following way:
If Ui , i ∈ N, are subsets for which `(Ui ) always is defined, then
∞
! ∞
[ X
` Ui ≤ `(Ui ).
i=1 i=1
Now assume that A = {ai | i ∈ N} is the range of a sequence. We call those sets
countable1 Choose some natural n > 1 and consider
1 1
Ui = (ai − i
, ai + i ).
n n
2
We have `(Ui ) = ni
. We therefore arrive at
∞
X 2 1
`(A) ≤ i
=2·
i=1
n n−1
using the geometric series (note, however, that the summation starts with i = 1 ).
This shows (with n going to infinity) that `(A) necessarily is 0.
The procedure motivates the following definition:
If for a set A ⊆ R and for every > 0 there exists a sequence of intervals
Ii = (ai , bi ) ⊆ R, such that
∞
[ ∞
X
A⊆ Ii and (bi − ai ) ≤ ,
i=1 i=1
(a) The fact that the unit interval [0, 1] has length 1 implies that it is not a
null set. As we know from 7.1.1 that countable sets are null sets, we see that
[0, 1] cannot be countable, it therefore cannot be the range of a sequence.
1
Some people prefer “at most countable” and treat the finite sets indpendently. For our
purposes, however, it is more convenient to call the finite sets countable as well.
7.1. LEBESGUE INTEGRAL 107
(b) It is quite surprising that the rational numbers Q are a countable set.
However, we already saw that in 3.2.4 (b), where we constructed a sequence
attaining every rational number as a value! Therefore, rational numbers are
a null set.
As a consequence, the set of irrational numbers cannot be a null set, because
otherwise [0, 1] were a null set, which it is not. This implies that in a certain
sense, there are much more irrational numbers than rational ones.
(c) There are also null sets which are not countable. A prominent example
is the Cantor set. This can be constructed by starting with [0, 1] = C0 ,
and once you have Cn which is a union of 2n intervals of length 1/3n ,
you construct Cn+1 by deleting the (open) middle third of every of these
previously remaining intervals.
The first few sets in this sequence are
C0 = [0, 1],
C1 = [0, 1/3] ∪ [2/3, 1],
C2 = [0, 1/9] ∪ [2/9, 1/3] ∪ [2/3, 7/9] ∪ [8/9, 1].
Then define the Cantor set to be the intersection of all these Cn ,
∞
\
C∞ = Cn .
n=0
There will be values σ(x0 ), σ(x1 ), . . . , σ(xn ) as well, but we do not make any
instructions concerning these. This has the advantage that sums of step functions
again are step functions.
108 KAPITEL 7. INTEGRATION
We already define the integral of the step function σ over the interval I as
Z b X n
σ(x)dx = ci · (xi − xi−1 ).
a i=1
This is motivated by the area of a rectangle being the product of its side lengths.
Note, however, that the ci may be negative, so that the interpretation of the
integral as an area is not fully justified - unless you speak about “signed” area.
Rb
If all ci are non negative, the integral a σ(x)dx indeed is the sum of the areas of
the rectangles with side length (xi − xi−1 ) and ci . This is the area of the region
between the x -axis and the graph of σ.
Then we define Z b Z b
f (x)dx = lim σn (x)dx.
a n→∞ a
We will explain later why this limit does not depend on the chosen sequence of
stepfunctions.
R Rb
We will also write I f (x)dx = a f (x)dx if I = [a, b].
f : I = [0, b] → R, f (x) = x.
b2
We expect the integral to be 2
, the area of a rectangular triangle with both legs
equal to b.
7.1. LEBESGUE INTEGRAL 109
We saw this function in Example 4.1.4, where we had defined it on the real line.
Now we restrict the domain to the unit interval. The constant function 1 on [0, 1]
is a step function which is nonnegative and coincides with f almost everywhere,
so we can use it to calculate
Z 1
f (x)dx = 1.
0
In fact, we could have prescribed arbitrary values for f at the rational numbers
without changing the integral, because the rational numbers are a null set.
Proof. We have to show that, if (σn )n and (τn )n are two sequences of non negative
step functions such that for almost all x ∈ [a, b] the sequences (σn (x))n and
(τn (x))n grow monotonically and converge to f (x),
Z b Z b
lim σn (x)dx = lim τn (x)dx.
n→∞ a n→∞ a
In order to achieve that we fix some natural number m and consider the difference
δ̃n (x) := τm (x) − σn (x), x ∈ [a, b].
We then define
δn (x) := max(δ̃n (x), 0),
such that δn is a step function with non negative values, and as m is fixed, we
get δ̃n+1 (x) ≤ δ̃n (x) for all x, so that δn (x) also decreases.
As σn (x) converges to f (x) which is at least τm (x) almost everywhere, we will
have limn→∞ δn (x) = 0 almost everywhere. We now invoke without proof, that
under this condition Z b
lim δn (x)dx = 0.
n→∞ a
This is intuitively plausible, but not really easy to show.
As a consequence,
Z b Z b Z b
τm (x)dx ≤ lim σn (x)dx + lim δn (x)dx = lim σn (x)dx.
a n→∞ a n→∞ n→∞ a
Now we may let m tend to infinity and see that
Z b Z b
lim τm (x)dx ≤ lim σn (x)dx.
m→∞ a n→∞ a
Exchanging the roles of τm and σn shows that “ ≥ ” also is true, hence both
limits coincide.
7.2. THE MAIN THEOREMS ON INTEGRATION 111
and call this the integral of f over the interval [a, b].
We denote by L(I) the set of all Lebesgue integrable functions.
Rb
(d) If f (x) ≥ 0 almost everywhere and a
f (x)dx = 0, then f (x) = 0 almost
everywhere.
(e) For every c ∈ (a, b), the restrictions of f to [a, c] and [c, b] are integrable
and Z b Z c Z b
f (x)dx = f (x)dx + f (x)dx.
a a c
Proof.
(a) This assertion is clear for functions in L+ (I) and λ ≥ 0, because given
sequences (σn )n , (τn )n of non negative step functions which calculate the
integrals of f and g respectively, the step functions σn + λτn will do so for
f + λg.
For general f and g the assertion can be deduced from this by making
several case distinctions for f + , f − , g + , g − .
(b) We treat the case of non negative functions only. If (σn )n , (τn )n are se-
quences of non negative step functions which calculate the integrals of f
and g respectively, we can copy the proof from 7.1.7 and see from g −f ≤ 0
almost everywhere that
Z b Z b
τm (x)dx ≤ lim σn (x)dx.
a n→∞ a
Rb
But this limit just is a
f (x)dx, and as m goes to infinity, the claim follows.
The last assertion follows by symmetry.
(c) We omit the proof.
(d) If (σn )n is a sequence of non-negative step functions calculating the Le-
Rb
besgue integral of f, a σn (x)dx > 0, if σn is non-zero. The monotonicity
implies that all σn have to be zero. As they converge to f almost every-
where, we have the assertion.
7.2. THE MAIN THEOREMS ON INTEGRATION 113
(e) It is enough to show this for non negative functions, in which case one can
just consider the case of step functions σn as in the definition. For every
n, Z b Z c Z b
σn (x)dx = σn (x)dx + σn (x)dx
a a c
and this identity carries over to the limit by 3.1.7.
This will be helpful for circumvening several case distinctions later on.
Proof. As f has a maximal value by 4.2.5 there is some C > 0 such that f +C is
positive. If we can show that this is integrable, then f = (f +C)−C is integrable,
as a constant function certainly is.
Therefore it suffices to consider the case of f (x) ≥ 0 for every x.
We construct a family of step functions σn ≥ 0 in the following manner. Fix a
natural number n. Consider the subintervals Ii = [a + (i − 1) · b−a2n
, a + i · b−a
2n
], i =
n
1, . . . , 2 . Let vi be the minimal value of f on this compact interval, which exists
again by 4.2.5. Then σn is defined to be the function taking the value vi on the
interval Ii r {a + i · b−a 2n
} and 0 at b.
As every interval for the construction of σn+1 is contained in one of the intervals
of the generation before, the minimal value will not be smaller than there, and
(σn )n is a monotonically increasing sequence of step functions.
Now let x ∈ [a, b) and some > 0 be given. As f is continuous there exists
some δ > 0 such that
|f (x) − f (y)| ≤ , if y ∈ (x − δ, x + δ) ∩ I.
1
As soon as n satisfies 2n
< δ, we therefore see that
|σn (x) − f (x)| ≤ .
This implies that f (x) = limn→∞ σn (x).
Rb
As lastly the sequence ( a σn (x)dx)n is monotonically increasing and obviously
bounded by (b − a) · max(f (I)), we get convergence of this sequence by 3.2.8.
Therefore all we have demanded for f to be integrable is satisfied.
114 KAPITEL 7. INTEGRATION
The crucial point is that in 7.2.8 the interval [a, b] is contained in some larger
interval where f can be extended in a reasonable way. There really is something
to do in order to prove this equality if such an extension is not possible.
F (t)|t=b
t=a = F (b) − F (a).
where the functions are only fixed up to some additive constant, which can be
chosen as it does not change the derivative.
The author would prefer a notation like
Z x
f (t)dt,
to stress that the variable of F (x) is not the variable for the integration but the
upper bound in the integral. The missing lower bound still leaves place for the
additive constant.
xr+1
F (x) =
r+1
is an antiderivative of f.
7.3. CALCULATING SOME INTEGRALS 117
(a) In order to integrate ln(x) you can use f 0 (x) = 1, g(x) = ln(x). Together
with f (x) = x and g 0 (x) = x1 this gives for positive a, b
Z b x=b Z b
ln(x)dx = [x ln(x)] − 1dx = b ln(b) − a ln(a) − b + a.
a x=a a
The integral Z r √
A= r2 − x2 dx
−r
7.3. CALCULATING SOME INTEGRALS 119
x = f (t) = r cos(t), 0 ≤ t ≤ π.
The above formula then reads as (note the changed roles of x and t ;-))
Z πp Z π
A=− r2 − r2 cos2 (t) · (−r sin(t))dt = r2 sin2 (t)dt.
0 0
1
sin2 (x) = (1 − cos(2x)),
2
hence π
π
r2 r2
Z
sin(2t) π
A= · (1 − cos(2t)) dt = t− = r2 .
2 0 2 2 0 2
As we have
cosh(t)2 − sinh(t)2 = 1 and cosh(t) > 0,
we find q
cosh(t) = 1 + sinh2 (t).
We therefore substitute x = sinh(t) and choose α, β such that
sinh(α) = a, sinh(β) = b,
leading to
Z b Z β Z β
1 1
√ dx = q cosh(t)dt = 1dt = β − α.
a 1 + x2 α 1 + sinh2 (t) α
f 0 (x) = u(x),
and
R x we just learned how to solve this on an interval, using the indefinite integral
u(t)dt.
It might therefore be a good place to say something more on at least certain
classes of differential equations we can deal with right now.
ẍ(t) = g.
This leads to
ẋ(t) = g · t + v0 , v0 a constant,
and this again leads to
1
x(t) = gt2 + v0 · t + x0 , x0 a constant.
2
Of course for a specific movement there should be no choices of the constants –
they should be prescribed from the circumstances. But then we see that x0 = x(0)
is the place, where the point is at time 0, and v0 = ẋ(0) is the speed at time 0.
And of course the movement will depend on where the point happens to be at
time 0 and whether you just let it fall or throw it into the air or smash it to the
ground.
Thus in addition to the differential equation itself we also will need certain “boun-
dary conditions” to get a solution for a specific problem.
so that the weight can oscillate freely without being hindered by whatever. . . only
gravitation plays its role. We describe the location of the weight by its angle φ
with respect to the vertical axis emanating from the upper end of the rope. The
angle is a function φ(t) with respect to time, and it turns out that its acceleration
is given by
g
φ̈(t) = sin(φ(t)).
l
This equation already is quite hard to solve, and it is much nicer to look at the
approximation for small angles, i.e. in the case that the initial displacement is
not too large. Then sin(φ(t)) ∼ φ(t), and we see solutions, namely
p p
a · cos( g/l · t) + b · sin( g/l · t), a, b ∈ R.
One can even show (but not we, at least not now!) that these are the only solu-
tions.
...
..... ......
...
......... ..
...... ....... .. Fg
.......
........ .
.. ........... .
......... ..
........ ..... ..... ..... ..... ......
. .
........... .....
.................
. ..
...
...
...
. .
. ............
. F
.................... .................................. ...
... −F
−F ... g
...
.........................................................................•...................................................................•.............................
0 x
The forces which contribute are the two tension forces at the ends of the red part
of the rope (in tangential direction) and the gravitational force. We decompose
the tension force at the right end into its vertical and horizontal components
(dashed blue lines). As the three forces have to add up to zero (otherwise we
would have some movement of this part of the rope) we see that the horizontal
part at the right end is the same as that on the left end (other direction) and
that the vertical part at the right end is the gravitational force. As the tension
122 KAPITEL 7. INTEGRATION
force at the right end is tangential to the curve, its slope is f 0 (x), and we get
that
f 0 (x) · F = Fg = density · (length of red curve).
Here, “density” is understood as the density with respect to length; the rope is
an ideal rope, homogeneous and without width;-)
As the force F at the left end is independent of x we see that f 0 (x) is propor-
tional to the length of the red curve, which depends on x. This is the arc length
of the curve
γ : [0, x] → R2 , γ(t) = (t, f (t)).
The velocity of this curve is γ 0 (t) = (1, f 0 (t)), and the absolute value of the speed
therefore p
kγ 0 (t)k = 1 + f 0 (t)2 .
Therefore the length of the red curve is – cf. 6.1.5 –
Z xp
1 + f 0 (t)2 dt.
0
We arrange the units (for length, weight) such that the above mentioned propor-
tionality becomes an equality, and get
Z xp
0
f (x) = 1 + f 0 (t)2 dt.
0
As the slope of the rope grows monotonically, we have f 00 (x) > 0 everywhere
such that we may divide by f 00 (x) :
We now remember that f 0 (0) = 0, because f has a local minimum there, and
hence f 00 (0) = 1, as it is positive with square 1. We therefore can write down the
Taylor series for f 0 at 0 , which is
∞ ∞
X
(n+1) n
X x2n+1
Tf 0 (x) = f (0)x /n! = ,
n=0 n=1
(2n + 1)!
because the even derivatives of f 0 vanish at 0 and the odd derivatives are 1.
Therefore, cf. 5.4.8, (the Taylor series of) f 0 (x) = sinh(x) and f (x) = cosh(x)+c,
where c depends on the heigth where the nails holding the rope are hit into the
wall.
* * * * *
We now study three types of differential equations of first order, i.e. only involving
the unknown function and its first derivative. These types are all of the shape
d u0 (x) !
H(u(x)) = = g(x),
dx h(u(x))
such that H(u(x)) is an antiderivative of g, which we can calculate by integra-
tion, such that Z x
−1
u(x) = H g(t)dt
x0
124 KAPITEL 7. INTEGRATION
solves the original equation. Note, that H is injective as h(u) never gets zero,
which means there are no sign changes for the values of h and therefore H is
monotonic.
This is the general strategy, and we now look at two more special cases.
u0 (x) = h(u(x)).
This is the case if, for instance, the physical conditions which are imposed
on the movement of a particle at position u in dependence on time x do
not change during the time of observation.
Then we have G(x) = x+const. and u(x) = H −1 (x+const.). There is also
an additive constant involved in the choice of H, and we should arrange
for the constants in such a way, that H −1 (x + const.) is defined at least at
x = x0 und such that the value there is our boundary value u0 – if possible.
No one can promise that such a choice always is possible, and in particular
it might not be possible to define u on all of I.
Sometimes one is particularly interested in stationary solutions, which are
just constant functions and are possible if h(u0 ) = 0. Then it will be
interesting whether this solution is stable, i.e. small changes will lead to a
new solution which approaches u0 , or unstable, i.e. a small change of the
conditions leads to a new solution moving away from u0 . But as you see,
in this case the old assumption that h never gets zero is no longer valid, so
that the original strategy will not work unchanged.
(b) The case h(u) = u, the so called homogeneous linear differential equation
of first order:
u0 (x) = g(x) · u(x).
We already dealt with this type of equation in 6.4.4 and saw the specific
solution
u(x) = exp(G(x)), where G0 = g.
We can now calculate such a G for every continuous g on an interval
I, using integration. We now take a second solution ũ of the equation
and compare it to the first solution u by dividing ũ by u, i.e. looking at
v(x) = ũ(x) exp(−G(x)).
Differentiation as well as application of the product rule and the chain rule
lead to
x 7→ c · u(x), c ∈ R constant,
Note that we are no longer in the separable case and that h has a different
meaning here. The function h is called the inhomogeneity of this equation.
As we can already solve the equation in case the inhomogeneity should be zero
we could try to start with a solution of the homogeneous equation u0 = gu
and modify this in order to get a solution of the inhomogeneous equation we
are interested in. The idea is that the derivative of a product is a sum, and we
now try to arrange for suitable factors. The attempt we make is the variation of
constants and uses a function
Here the first summand already is g(x) · u(x), so we should try to solve for
v 0 = g(x) · v(x),
and this shows that for a fixed solution us (the index s standing for “special”)
the functions
us (x) + a · exp(G(x)), a ∈ R,
are all solutions of the inhomogeneous linear differential equation. Again, for every
boundary condition u(x0 ) = u0 there is a unique value of a = (u0 − us (x0 )) ·
exp(−G(x0 )) giving the right solution.
We will see an example of this in 7.4.8.
λv 0 = u0 · v 1−λ = g · v + h · v λα+1−λ .
1
If λα + 1 − λ = 0, i.e. λ = 1−α , this is an inhomogeneous linear differential
equation for v (after dividing by λ) .
As now we have to transform back to u we will often have to assume that v only
takes positive values, such that u = v λ is defined.
We now solve such an equation in an example.
As now we want v 2 = u, we also get the boundary condition v(1) = 1, and this
forces k = 1, so that our solution will be
p
u(x) = x · 2 ln(x) + 1.
But as we want a real square root, we have to demand that the expression in the
square root is non-negative, leading to x ≥ √1e , so that the solution cannot be
defined for all positive real numbers.
In the general task to solve differential equations this already indicates that we
will have to find a suitable (maximal. . . ) domain where the equation can be
solved, and this will also depend on the boundary conditions.
interval I (or even some other subset of the reals) is a step function on R which
has value zero outside I. In particular, such a step function is zero outside some
compact interval, as it has finitely many values on finitely many intervals of finite
length.
We can then word by word transfer the definition of the Lebesgue integral to the
case of unbounded intervals.
The task now still is to evaluate such integrals, as we can no longer use the
Fundamental Theorem 7.2.9, having no end points of the interval where we can
evaluate the antiderivative. The natural idea of course is to approximate the
domain by compact intervals. The technical tool for doing this is the following
theorem which we state in a special case only.
This gets slightly more interesting when the integrand is somewhat more intricate,
e.g.
Z ∞ Z b Z b
−x −x −x b −x
xe dx = lim xe dx = lim (−xe ) 0 + e dx = 1,
0 b→∞ 0 b→∞ 0
−x
because limb→∞ xe = 0 due to 6.3.11.
R∞
Integrals of the shape 0 f (x)e−x dx are special values of the Laplace transform
of f. Laplace transforms are a tool for solving differential equations and will be
introduced in the AM2-course as close relatives to Fourier transforms.
The first two integrals evaluated above
Pd can ibe extended inductively, showing
finally that for a polynomial f (x) = i=0 ai x
Z d
X
−x
f (x)e dx = ai · i!.
i=0
For instance,
1 1 1
√
Z Z
1 1
√ dx = lim √ dx = lim 2 x = 2.
0 x j→∞ 1/j x j→∞
1/j
130 KAPITEL 7. INTEGRATION
and therefore
kπ k
| sin(x)|
Z
2X1
dx ≥ ,
0 x π j=1 j
1
because |f (x + π)| ≤ |f (x)|, as | sin(x + π)| = | sin(x)| and x+π
< x1 .
Due to the signs of the sine function, we have
Z kπ k
X Z jπ
j+1
f (x)dx = (−1) |f (x)|dx
0 j=1 (j−1)π
5.1.10 Rβ R (k+1)π
≤ | kπ
f (x)dx| + | kπ
f (x)dx|
R (k+1)π 2
≤ 2| kπ
f (x)dx| ≤ kπ
≤ β2 ,
exists.
R 2kπ R∞
However, although limk→∞ 0
sin(x)dx = 0, the improper integral 0
sin(x)dx
does not exist.
There are other obvious generalizations of the existence of the improper integral,
e.g. replacing sin by some piecewise continuous function s : [0, ∞) → R for
which a positive h exists with
s(x + h) = −s(x)
Looking more closely at partial fractions, there are also summands of the type
ax+b
(x2 +cx+d)n
, were the polynomial function x2 + cx + d does not have a real root.
This is the case if and only if 4d − c2 ≥ 0. Note, that therefore x2 + cx + d is
132 KAPITEL 7. INTEGRATION
always non negative, which allows us to take its square root in the real numbers.
We use this now after completing the square:
2
1
1 1 1 x + c
(x2 + cx + d) = ((x + c)2 + d − c2 ) = (d − c2 ) · q 2 + 1 .
2 4 4 1 2
d − 4c
a d b− ac 1
= 2
· dx
ln(x2 + cx + d) + 2
d− 41 c2
· 1c
!2
√x+ 2
1 c2
+1
d− 4
h i
d a √2b−ac
= dx 2
2
ln(x + cx + d) + 4d−c2
arctan( √2x+c
4d−c2
) .
As the degree of the numerator of f is smaller than that of the denominator, our
decomposition will be of the shape
A B Cx + D
+ 2
+ 2 ,
x + 1 (x + 1) x +2
Comparing the coefficients of the numerator here with those of the numerator of
f leads to a system of four equations in the variables A, B, C, D :
A +C =3
A +B +2C +D =8
2A +C +2D =6
2A +2B +D =7
One now can solve for this equations by using Gaußian elimination. For instance,
subtracting two times the first equation from the third gives C = 2D, hence
A = 3 − 2D. Plugging this in equations 3 and 4 results in
B +3D = 5
2B −3D = 1
which after addition gives 3B = 6, and then successively
B = 2, D = 1, C = 2, A = 1.
Therefore
1 2 2x + 1
f= + 2
+ 2 ,
x + 1 (x + 1) x +2
and this has antiderivative
−2 1 x
ln |x + 1| + + ln(x2 + 2) + √ arctan( √ ) + const.
x+1 2 2
To be honest, this derivation of an antiderivative does not at all use the theory
of integrals and could have been performed much earlier. But now we might
have some more motivation to calculate antiderivatives because they have a new
meaning as an indefinite integral.
Indices
Bolzano-Weierstraß 3.2.5
Cauchy Convolution Formula 5.1.7
Chain Rule 6.2.3
Comparison test 5.2.1
Existence of maxima 4.2.5
Functional Equation 5.1.8
Fundamental theorem of differential and integral calculus 7.2.7,7.2.8
Intermediate Value Theorem 4.2.7
Leibniz Criterion 5.1.10
L’Hôpital’s rule 6.3.9
Mean value theorem 6.3.4
Mean value theorem for integral calculus 7.2.5
Monotonicity Criterion 3.2.8
Ratio Test 5.2.3
Rolle’s Theorem 6.3.3
Root Test 5.2.5
135
136 KAPITEL 8. INDICES
8.3 Terminology