Fourthintroduction To Transcendental
Fourthintroduction To Transcendental
Numbers
i
Contents
Contents iii
1 Introduction 1
1.1 Recap of Some Algebraic Concepts . . . . . . . . . . . . . . . . 1
1.1.1 Algebraic Numbers . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Algebraic Integers . . . . . . . . . . . . . . . . . . . . . 8
1.1.3 Transcendental Numbers . . . . . . . . . . . . . . . . . 12
1.2 Some Historical Notes . . . . . . . . . . . . . . . . . . . . . . . 14
2 Transcendence Criteria 17
2.1 Irrationality Criterion . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.1 A Proof of the Irrationality of e . . . . . . . . . . . . . . 24
2.2 Liouville’s Approximation Theorem . . . . . . . . . . . . . . . 25
2.2.1 Liouville Numbers . . . . . . . . . . . . . . . . . . . . . 31
2.2.2 Generalizations of Liouville’s Approximation Theorem 36
3 Transcendence of e 39
3.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 The Original Proof of Hermite . . . . . . . . . . . . . . . . . . 45
iii
Chapter 1
Introduction
1
1. Introduction
deg(irrα,Q ) = n ⇐⇒ [Q(α) : Q] = n,
i.e. the degree of the simple field extension Q(α)/Q is finite and equal to n.
In this case, the extension Q(α)/Q has for Q-basis the set
n o
1, α, α2 , . . . , αn−1
2
1.1. Recap of Some Algebraic Concepts
i.e. any x ∈ Q(α) can be written as a unique linear combination of the elements
1, α, α2 , . . . , αn−1 over Q, that is:
n −1
x= ∑ λi αi for some unique λi ∈ Q (i = 0, 1, . . . , n − 1)
i =0
deg(α) = 1 ⇐⇒ α ∈ Q
In particular, the degree of a real algebraic number is strictly more than 1 if and only
if α is an irrational number.
i.e. a number is algebraic if and only if its real and imaginary parts are algebraic
numbers.
3
1. Introduction
irrα,Q = irrα,Q
irrRe(α),Q divides g in Q[ X ].
4
1.1. Recap of Some Algebraic Concepts
E = Q( α1 , . . . , α n )
E = Q[ α1 , . . . , α n ]
α1 , α2 ∈ { α1 , α2 }
We know that any element σ ∈ Gal ( f ) permutes the roots of f , which are by
definition the algebraic conjugates of α, and in fact any σ ∈ Gal ( f ) is totally
determined by its values at the algebraic conjugates of α. This result comes
from the existence of a well-known injective group homomorphism Φ from
5
1. Introduction
the Galois group Gal ( f ) to the set of all permutations of the set of the roots
of f , namely R E ( f ) := { β ∈ E | f ( β) = 0}, given by:
Φ : Gal ( f ) ,−
→ SRE ( f )
σ 7→ σ
RE ( f )
g = g( X ) = bn X n + bn−1 X n−1 + . . . + b1 X + b0 ∈ R[ X ],
we have that:
therefore all bi ’s belong to the fixed field EGal (E/Q) of Gal ( f ) = Gal ( E/Q).
Since E is the splitting field of a polynomial in Q[ X ] and since char (Q) = 0,
we have that the field extension E/Q is Galois, which is equivalent to:
EGal (E/Q) = Q
C = Q t (R \ Q) t (C \ R),
and since rational numbers are simply algebraic numbers of degree 1, the subfield A
of C is divided into three non-overlapping sets:
Q and (R \ Q) ∩ A and (C \ R) ∩ A
6
1.1. Recap of Some Algebraic Concepts
We now give a result about the cardinality of the set of algebraic numbers.
Proof We only give the outline of the proof. We first recall that any polyno-
mial with coefficients in a given field has a finite number of roots which is at
most its degree. Now, since the set Q of all rational numbers is countable,
the set Q[ X ] of all polynomials with rational coefficients is also countable.
Therefore, since the set A of all algebraic numbers is the set of all roots of
a countable number of polynomials, each with a finite number of roots, we
have the desired result.
There is another property of algebraic numbers using polynomials with
integral coefficients, instead of rational coefficients:
∃ a non-constant polynomial g ∈ Z[ X ]
α ∈ A ⇐⇒
such that g(α) = 0 and with positive leading coefficient
Proof One direction is obvious, because any polynomial with integral coeffi-
cients is certainly a polynomial with rational coefficients.
For the other direction, let α ∈ A be an algebraic number. That is, there
exists a polynomial f = f ( X ) = an X n + . . . + a1 X + a0 with n ∈ N≥1 and
an , . . . , a1 , a0 ∈ Q and an 6= 0 and such that f (α) = 0. We set ai := pi /qi
7
1. Introduction
8
1.1. Recap of Some Algebraic Concepts
√
Proof Assume α := 2/3 is an algebraic integer. Then there exists a monic
polynomial g = g( X ) = X n + bn−1 X n−1 + . . . + b1 X + b0 ∈ Z[ X ] which has
α as root. So:
√ !n √ ! n −1
2 2
g(α) = + bn − 1 + . . . + b0 = 0
3 3
Since 3 divides 0 in Z, each sum above (the one for i even and the one for i
odd) must be divisible by 3. In particular, because each summand containing
bi , i 6= n, has a factor of 3, the number 3 must divide the summand containing
n −1 n
bn = 1. This tells us that 3 divides 2 2 in Z if n is odd, and 3 divides 2 2 in
Z if n√is even. In either case, this is false and hence we can conclude that
α := 2/3 is not an algebraic integer.
We have to stress out
√ that algebraic integers are not always (ordinary) integers.
For example, i or 2 are algebraic integers
√ which are not ordinary integers.
Indeed, we have g1 (i ) = 0 and g2 ( 2) = 0 with g1 = g1 ( X ) = X 2 + 1
and g2 = g2 ( X ) = X 2 − 2 being two non-constant, monic polynomials with
integral coefficients.
In fact, we can specify the nature of real algebraic integers.
9
1. Introduction
10
1.1. Recap of Some Algebraic Concepts
bd · g = ac · ϕ∗ ψ∗
bd ∼ ac in Z,
11
1. Introduction
Proof We only outline the proof which uses some linear algebra. Let
E := Q( an , an−1 , . . . , a1 , a0 ) be the smallest field containing Q and all the
coefficients of h = h( X ) = an X n + an−1 X n−1 + . . . + a1 X + a0 . Since all co-
efficients of h are algebraic elements over Q, then the field extension E/Q
is a finite extension. Since the vector space formed by the powers of α is
finite-dimensional over E, it is also finite-dimensional over Q. That is, some
linear combination of the powers of α with rational coefficients vanishes, so
α is a root of a polynomial with rational coefficients.
12
1.1. Recap of Some Algebraic Concepts
We have seen examples of elements in the first two sets, but a priori, we do
not know yet whether the set of real transcendental numbers is empty or not
(it is not!)
One can even wonder if the set of non-real transcendental numbers is empty
or not. To partially answer this question, we recall Proposition 1.11, which
says that a complex number is an algebraic number if and only if its real and
imaginary parts are both algebraic numbers. By taking the contraposite, we
obtain:
Proof Assume the contrary, i.e. α1 + α2 and α1 · α2 are both algebraic numbers.
Then the polynomial
h = h( X ) = ( X − α1 )( X − α2 ) = X 2 − (α1 + α2 ) X + α1 α2 ∈ A[ X ]
13
1. Introduction
If yes, can one give explicit examples of real transcendental numbers? And
even more difficult: given a specific (necessarily irrational) real number, like
e or π, how can you determine whether this number is transcendental or not?
We will answer to all these questions in this report, but we start by giving a
brief historical outlook on the subject of the transcendental numbers.
14
1.2. Some Historical Notes
Hermite’s proof and by making use of the famous Euler equation linking
the old constant π to the modern constant e, namely eiπ + 1 = 0. The
proofs of the transcendence of e and π were considered among the greatest
achievements of 19th century mathematics. The transcendence of π especially
is a fact of great historical, as well as intrinsic, interest.
Indeed, one of the classical problems of Greek mathematics was to construct,
with compass and straightedge alone, a square whose area is that of a circle
√of
radius 1. This requires the construction of a line segment whose length is π,
which can be accomplished if a line segment of length π is constructible. The
Greeks were totally unable to decide whether such a line segment could be
constructed, and even the full resources of modern mathematics were unable
to settle this question until 1882, when π was proven to be transcendental
by Lindemann. Since the length of any line segment that can be constructed
√
with straightedge and compass can be written in terms of +, ×, −, ÷, and ,
and is therefore algebraic, this proves that a line segment of length π cannot
be constructed.
A totally different approach to the existence of transcendental numbers
was taken by Cantor who proved with his famous diagonal’s argument
(which dates back to 1874) that the real numbers are uncountable. Since
the real numbers are the union of the real algebraic numbers and the real
transcendental numbers, both of these sets cannot be countable. This implies
that the real transcendental numbers must be uncountable, since the real
algebraic numbers are countable (see Proposition 1.15). Cantor thus showed
in a non-constructive manner that almost all real numbers are transcendental,
without giving any explicit examples.
From the end of the 19th century up until the 1930’s, not much new impetus
has been given to the subject of transcendental numbers: we can only mention
a dozen of new and simpler proofs of the transcendence of e and π which rely
more and more on the property of the exponential function, and therefore
less succeptible of being generalized.
But in 1934 new impulse was given to the subject, when Hilbert’s 7th problem
(posed in 1900) was solved, namely:
15
1. Introduction
Among the important results of the remaining of the 20th century, we can
cite the classification of transcendental numbers by Mahler (1932) and later
by Koksma (1939); Schneider’s theorem on elliptic functions and Abelian
integrals (1941) and the Thue-Siegel-Roth Theorem on the approximation of
algebraic numbers by rationals (1955), which generalizes Liouville’s original
approximation theorem.
Open problems in the subject of transcendental numbers are very interesting
and surprisingly easy to state: they form a list of given numbers which have
yet to be proven to be either transcendental or algebraic.
For example, most sums, products, powers, etc. of √
the number π and the
e e 2 2
number e, e.g. eπ, e + π, π − e, π/e, π , e , π , π , eπ , are not known to
π
Here, b x c represents the floor function. The number γ has not been proved
algebraic or transcendental. In fact, it is not even known whether γ is
irrational. Of course, some progress have been made, but a definite answer
has still not been given.
In the same vein, the Riemann zeta function at odd integers strictly greater
than 3, namely ζ (5), ζ (7), . . ., has not even been proven to be irrational.
The subject of transcendental numbers is a branch of number theory that
investigates transcendental numbers, in both qualitative and quantitative
ways. As yet, there seems to be no aspect to the study of transcendental
numbers which could be described as a general theory. However, there
are some methods which are quite powerful and useful in other domain
of mathematics. For example, many approximation theorems used in the
study of transcendental numbers come in handy in the study of diophantine
equations.
16
Chapter 2
Transcendence Criteria
Usually the numbers for which we do not know of their transcendence arise
from analysis and are given by infinite processes, involving limits, infinite
series, infinite products or integrals. So it is very difficult in general to deter-
mine whether those numbers are even irrational, let alone transcendental, or
not.
Since being irrational is a necessary condition for a real number to be tran-
scendental, we start by looking at some irrationality criteria, hoping that one
of them can be modified into a transcendence criterion.
There exists a well-known criterion to determine whether a real and (for
simplification) non-negative number x is irrational or not, using its expansion
in an integer basis b ∈ N≥2 (e.g. in binary or decimal expansion), namely:
dj
x = bxc + ∑ j
= : [ s · d1 d2 d3 . . . ] b
j≥1 b
[ s · d 1 d 2 d 3 . . . d r d r +1 d r +2 . . . d r + p d r +1 d r +2 . . . d r + p d r +1 d r +2 . . . d r + p d r +1 . . . ] b
17
2. Transcendence Criteria
There is also another well-known criterion for irrationality using the contin-
ued fraction expansion. Any real and (for simplification) non-negative x can
be uniquely written as a simple continued fraction of the form
1
x = a0 + = : [ a0 ; a1 , a2 , a3 , . . . ]
1
a1 +
1
a2 +
1
a3 +
.
a4 + . .
p
|θ − | < ε (2.1)
q
p
The problem is that, as we try to make q closer and closer to θ, we may
have to use larger and larger integers p and q. So, the reasonable question
to ask here is how well can we approximate θ by rationals without too large
denominators.
p
Trivially, every real number θ can be approximated by a rational number q
1
with a given denominator q ≥ 1 with an error not exceeding 2q .
18
2.1. Irrationality Criterion
Let
indeed q ∈ Z>0 be any positive integer. Note that the closed interval
qθ − 2 , qθ + 12 has length 1 and, therefore, contains at least one integer.
1
This estimate
p 1
|θ − | ≤ (2.2)
q 2q
certainly implies (2.1). Indeed, for any ε > 0, just take q ∈ Z>0 large enough
1
so that 2q < ε and you get:
p 1
|θ − | ≤ <ε
q 2q
19
2. Transcendence Criteria
This way of distinguishing rational from irrational numbers using rational ap-
proximations by specifying the order of the approximation error is formalised
in the next proposition:
Proposition 2.1 (Criterion for irrationality / rationality) Let θ ∈ R be a real
number. Then the following conditions are equivalent:
(i) θ 6∈ Q
(ii) For any ε > 0 there exists a rational number p/q ∈ Q (with p, q ∈ Z, q > 0
and gcd( p, q) ∼ 1) such that
p ε
0 < |θ − |<
q q
i.e. θ is said to be well approximable by rational numbers.
p
(The inequality > 0 guarantees that we consider rationals q 6= θ)
(iii) (Asymptotic Dirichlet’s Theorem or Dirichlet’s Theorem on Diophantine
Equations) There exist infinitely many rational numbers p/q ∈ Q (with
p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1) such that
p 1
|θ − | < 2
q q
p
(We do not need to specify here that q 6= θ, since we have infinitely many of
them.)
(iv) (Uniform Dirichlet’s Theorem) For any real number Q > 1, there exists a
rational number p/q ∈ Q (with p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1) such
that 1 ≤ q < Q and with the property that
p 1
0 < |θ − |<
q qQ
Proof (ii ) ⇒ (i ): This result has already been proven, because we have
explained the contraposite ¬(i ) =⇒ ¬(ii ) (see Equation (2.3)), namely:
By the way the implication (ii ) ⇒ (i ) is actually the useful part of the
irrationality/rationality criterion, and paradoxically also the easiest one to
prove! The other easy implications are (iv) ⇒ (iii ) ⇒ (ii ) ⇒ (i ).
Indeed, let us prove that (iv) =⇒ (iii ):
Let Q > 1 be any real number. Then by the Uniform Dirichlet’s Theo-
rem, there exists a rational number p/q ∈ Q (with p, q ∈ Z, q > 0 and
20
2.1. Irrationality Criterion
1 1
gcd( p, q) ∼ 1) such that 1 ≤ q < Q (which implies that 1 ≥ q > Q) and such
that
p 1 1
0 < |θ − | < < 2.
q qQ q
p p
Let Q0 be any real number exceeding |θ − q |−1 (which implies that Q10 < |θ − q |).
A second application of Uniform Dirichlet’s Theorem shows that there exists
a rational number p0 /q0 ∈ Q (with p0 , q0 ∈ Z, q0 > 0 and gcd( p0 , q0 ) ∼ 1) such
that 1 ≤ q0 < Q0 (which implies that 1 ≥ q10 > Q10 ) and such that
p
p0 1 |θ − q | p
0 < |θ − 0 | < 0 0 < 0
≤ |θ − |
q qQ q q
p0 p
Thus, necessarily, we get: q0 6= q .
Furthermore, we have that:
p0 1 1
|θ − 0
|< 0 0 < 0 2
q qQ (q )
By iterating this process, we obtain a sequence { qii }i∞=1 of distinct ratio-
p
nal numbers (with pi , qi ∈ Z, qi > 0 and gcd( pi , qi ) ∼ 1) such that for all
i = 1, 2, . . .
pi p i −1 p1
| θ − q i | < | θ − q i −1 | < . . . < | θ − q 1 | ,
and
0 < | θ − pi | < 1 2
qi (q ) i
Let us now prove that (iii ) =⇒ (ii ): Since there exist infinitely many rational
numbers pi /qi ∈ Q (with pi , qi ∈ Z, qi > 0 and gcd( pi , qi ) ∼ 1) such that
pi 1
|θ − |< ,
qi ( q i )2
then the set
qi ∈ Z>0 | i = 1, 2, . . .
is infinite.
Otherwise, if there were only finitely many such q1 , q2 , . . . , q M ∈ Z>0 ,
then, since there can only exist for each qi (i = 1, 2, . . . , M) finitely many
(i ) (i ) (i )
p1 , . . . pri ∈ Z with gcd( p j , qi ) ∼ 1 (j = 1, . . . , ri ) which satisfy
(i )
pj 1
|θ − |< ,
qi ( q i )2
21
2. Transcendence Criteria
(i )
p
there would be only finitely many rationals qji (i =, 1 . . . , M and j = 1, . . . , ri )
which satisfy the inequality in (iii ), which is a contradiction.
For that reason, since {qi }i∞=1 is an infinite sequence of positive integers, we
have (by considering a subsequence if necessary): qi → ∞, or equivalently
qi → 0, as i → ∞.
1
p i0 1 1 1 ε
0 < |θ − |< 2
= · <
q i0 ( q i0 ) q i0 q i0 q i0
By the pigeonhole principle, since we have N+1 elements and only N subin-
tervals, one of these N subintervals, say Ij0 , contains at least two elements
of E. Apart from 0 and 1, all elements {qθ } in E with 1 ≤ q ≤ N − 1 are
j j +1
irrational, hence belong to the union of the open intervals ( N , N ) with
0 ≤ j ≤ N − 1.
Case 1: If j0 = N − 1, then the interval
1
Ij0 = IN −1 = 1 − , 1
N
But we have for the fractional part 0 ≤ {q∗ θ } < 1, implying: 1 − {q∗ θ } > 0.
Moreover, 1 and {q θ } lie in 1 − N1 , 1 with {q∗ θ } ∈ 1 − N1 , 1 , hence,
∗
22
2.1. Irrationality Criterion
since Q ≤ N:
1 1
0 < p∗ − q∗ θ = 1 − {q∗ θ } < ≤ .
N Q
So by dividing by q∗ > 0, we certainly have:
p∗ 1
0 < |θ − ∗
| < ∗ with 1 ≤ q∗ < Q,
q q Q
as desired.
Case 2: If j0 6= N − 1, we have that 0 ≤ j0 ≤ N − 2 and Ij0 contains two
elements {q1 θ } and {q2 θ } with 0 ≤ q1 < q2 ≤ N − 1. We have, since {q2 θ } is
j j +1
an irrational number lying in the open interval ( N0 , 0N ), that:
1
|{q2 θ } − {q1 θ }| < (2.4)
N
Now set
q : = q2 − q1 , p : = b q2 θ c − b q1 θ c .
Then we have 0 < q = q2 − q1 ≤ N − 1 < Q and
|qθ − p| = |q2 θ − q1 θ − bq2 θ c + bq1 θ c| = |{q2 θ } − {q1 θ }|
So we have from (2.4) and since Q ≤ N:
1 1
0 < |qθ − p| < ≤ ,
N Q
which implies that
p 1
0 < |θ − |< with 1 ≤ q < Q,
q qQ
as desired.
In both cases, the coprimality of the numerator and denominator of the
rational number p/q is obtained easily by dividing through by gcd( p, q).
A refined version of the Asymptotic Dirichlet’s Theorem (which dates back
to 1842) was found by Adolf Hurwitz in 1891. One can prove it using either
continued fractions or Farey sequences. We do not give a proof here, only
the statement:
Proposition 2.2 (Hurwitz’s Theorem) Let θ be a real number. Then the following
conditions are equivalent:
(a) θ is irrational
(b) There exist infinitely many rational p/q ∈ Q (with p, q ∈ Z, q > 0 and
gcd( p, q) ∼ 1) such that
p 1
0 < |θ − |< √
q 5q2
23
2. Transcendence Criteria
Remark 2.3 Of course Condition (b) in Proposition 2.2 looks stronger than con-
1
dition (iii ) in the irrationality criterion (since √5q 2
≤ q12 ), so we certainly have
(b) =⇒ (iii ) =⇒ θ ∈ R \ Q. But (b) is in fact equivalent to (iii ), and that is what
Hurwitz proved, by proving that ( a) =⇒ (b). √
Moreover, the estimate in (b) with the constant 1/ 5 is optimal, in the sense that √
it is the best possible approximation we can have for the golden ratio ϕ := 1+2 5 .
Indeed, one can prove that for any real number c < √15 , there are only finitely many
p
rationals p/q (with p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1) such that | ϕ − q | < qc2 .
is well-known to be irrational. One can find many proofs of this fact, even
very short ones. We will present a different proof of the irrationality of e
using the irrationality criterion.
p c p
|e − | > ⇐⇒ q · |e − | > c
q q q
1
0< <c
n
p∗
This real number is in fact a rational number of the form q∗ (with p∗ , q∗ ∈ Z,
p∗
q∗ > 0 and gcd( p∗ , q∗ ) ∼ 1) satisfying q∗ 6= e. To see this, just write the n+1
summands with the same denominator n!, and then set q∗ := n! and p∗ the
24
2.2. Liouville’s Approximation Theorem
25
2. Transcendence Criteria
and
for any real number Q > 1
∃ g = g( X ) := qX − p ∈ Z[ X ]
(iv) ⇐⇒ with degree 1 and even with leading coefficient 1 ≤ q < Q
which is Z -primitive (i.e. gcd( p, q) ∼ 1)
and such that 0 < | g(θ )| < Q1
By considering polynomials not just of degree 1, but of any degree, one can
modify the criterion for a number to be rational into a criterion for a number
to be algebraic (of any degree). This is exactly what Liouville did, as well as
later mathematicians who generalized Liouville’s result.
Liouville’s criterion essentially takes up the idea that not only rational num-
bers, but also algebraic numbers in general cannot be very well approximated
by rational numbers. So if a number can be very well approximated by
rational numbers, then it must be transcendental. The exact meaning of ”very
well approximated” is explained in the next proposition called ”Liouville’s
Approximation Theorem”.
Liouville’s criterion is unfortunately only a necessary condition for a number
to be algebraic, and hence by taking the contrapositive, Liouville’s criterion
can be rewritten as a sufficient condition for a number to be transcendental.
But his criterion was not strong enough to be necessary too, and indeed it fails
to detect the transcendence of e, for example. However, it did provide a large
class of real transcendental numbers. Liouville numbers have the privilege of
a tight approximation by rational numbers. This fact is against our intuitive
expectations, because it shows that in some respects real transcendental
numbers are nearer to rational numbers than algebraic irrational numbers.
Here is the classical version of Liouville’s criterion using an asymptotic
estimate:
holds for all rational numbers p/q ∈ Q \ {α} (with p, q ∈ Z, q > 0, gcd( p, q) ∼ 1)
which are different from α.
26
2.2. Liouville’s Approximation Theorem
criterion, there exists ε 0 > 0 such that for any rational numbers p/q
p
(with p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1), we have: |α − q | = 0 or
|α − qp | ≥ εq0 . So for any rational number p/q (with p, q ∈ Z, q > 0 and
gcd( p, q) ∼ 1) which are different from α, we must have:
p ε0
|α − | ≥
q q
27
2. Transcendence Criteria
p
• |α − q | < 1
So we obtain:
| an pn + an−1 pn−1 q + . . . + a1 pqn−1 + a0 qn |
p 1
|g |= ≥ n
q qn q
1 p
≤ | g0 (ξ )| · |α − | (2.5)
qn q
28
2.2. Liouville’s Approximation Theorem
evaluated at ξ, we get:
| g0 (ξ )| ≤ n · A · n · (1 + |α|)n−1 = n2 · A · (1 + |α|)n−1
| g0 (ξ )| ≤ K (2.6)
p
By multiplying both sides of (2.6) by |α − q | > 0 (because α is
irrational) and by considering (2.5), we get the inequalities:
1 p p
n
≤ | g0 (ξ )| · |α − | ≤ K · |α − |
q q q
which implies that
p 1 1 1 1
|α − | ≥ · n > · n
q K q 2K q
1
By taking c := 2K > 0, we get the desired result.
3. The case where α ∈ C \ R is a non-real algebraic number.
Then α is necessarily of degree n ≥ 2 (otherwise it is rational by Propo-
sition 1.7 and hence real, which is a contradiction). So by Proposition
1.13, since α is an algebraic number of degree n, then Re(α) is also an
algebraic number with degree m with 1 ≤ m ≤ n.
Since Re(α) is a real algebraic number of degree m ≥ 1, we can apply
the proven cases 1 and 2 of Liouville’s Approximation Theorem to get
a positive constant c̃ = c̃(Re(α)) > 0 depending only on Re(α) such
that the inequality
p c̃
|Re(α) − | > m (2.7)
q q
29
2. Transcendence Criteria
p p
|α − | = |(Re(α) − ) + iIm(α)|
q q
s
p 2
= Re(α) − + (Im(α))2
q
p
≥ | Re(α) − |
q
c̃
> m
q
c̃
≥ n
q
p p
|α − | = |(Re(α) − ) + iIm(α)|
q q
= |iIm(α)|
= |Im(α)|
Im(α)
> ·1
2
Im(α)
2
≥
qn
p c
|α − | > n ,
q q
as wanted.
30
2.2. Liouville’s Approximation Theorem
i.e.
α = 0.1100010000000000000000010000000000000000000000000000000000 y
0000000000000000000000000000000000000000000000000000000000001 . . .
Remark 2.5 The series defining α is convergent, since it is dominated by the geomet-
ric series ∑∞ 1
k =1 10k , hence α is a real number. Moreover, α is obviously an irrational
number, because its decimal expansion is not eventually periodic, because it has
arbitrarily long strings of 0’s since the 1’s appear at ever growing positions after the
decimal point.
p c
|α − | > n (2.8)
q q
holds for all rational numbers p/q ∈ Q (with p, q ∈ Z, q > 0, gcd( p, q) ∼ 1).
We do not need to require that p/q be different from α, since α is irrational.
Let i ∈ N be large enough so that
2
<c (2.9)
10(n+i)! · i
(This is possible, since c > 0 and the sequence { 10(n2+i)! · i }i∞=1 goes to 0 as i
goes to ∞).
Set j := n + i so that, with this new notation, we have
2
<c
10 j! · i
31
2. Transcendence Criteria
j
1
p := 10 · j!
∑ 10k!
k =1
1
j! 1 1
= 10 · ( + + . . . + j! )
101! 102! 10
= 10 j!−1! + 10 j!−2! + . . . + 10 j!− j!
= 10 j!−1! + 10 j!−2! + . . . + 1
so that
j
p 1 1 1 1
= 1! + 2! + . . . + j! = ∑ k!
q 10 10 10 k =1
10
We certainly have that p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1. So p/q is a relevant
rational number which satisfies inequality (2.8):
p c
|α − | > n
q q
1 1 1 1 10
1+ + 2 + 3 +... = 1
= <2
10 10 10 1 − 10 9
32
2.2. Liouville’s Approximation Theorem
So we have, since j := n + i:
p 1 2 2 2 2 1
|α − | < ( j+1)! · 2 = j! · ( j+1) < j! · j = j! · (n+i) = j! · i · j! · n
q 10 10 10 10 10 10
2
Now because q := 10 j! and since 10 j! · i
< c, we obtain:
p 1
|α − | < c · n (2.11)
q q
∞
1
α := ∑ bk!
k =1
with b being any integer ≥ 2 and with { ak }∞ k=1 being any sequence of integers
such that ak ∈ {0, 1, 2, . . . , b − 1} for all k and ak 6= 0 for infinitely many k, is
also a transcendental number.
In fact, we can construct in a similar manner infinitely many transcendental
numbers. They form a class of transcendental numbers, which we now call
Liouville numbers in honour of Liouville.
p 1
0 < α− < n.
q q
2
<1 (2.12)
10(n+i)! · i
33
2. Transcendence Criteria
34
2.2. Liouville’s Approximation Theorem
p 1 1 1
|α − | ≥ > n −1 ≥ n
q bq 2 q q
holds for all integers p and q with q > 0 and gcd( p, q) ∼ 1. Let r be a
positive integer for which 2r ≥ 1/c. Since α is a Liouville number, then for
the positive integer n + r, there are integers p∗ and q∗ with q∗ > 1 such that
p∗ 1 1 c
|α − | < ∗ n +r ≤ r ∗ n ≤ ∗ n
q∗ (q ) 2 (q ) (q )
Corollary 2.11 There are transcendental numbers which are not Liouville numbers.
Proof Since the set A of algebraic numbers is countable, the set of real
algebraic numbers in [0, 1] has certainly measure 0. Hence, the set of real
transcendental numbers in [0, 1] has measure infinite (equal to the power of
the continuum), like the measure of [0, 1]. Hence the set of real transcendental
numbers in [0, 1] cannot be the set of Liouville numbers in [0, 1], because
their measure differ (enormously!)
Finding a transcendental number which is not a Liouville number is not easy.
The numbers e and π are two such examples. One can prove for example that
e is not a Liouville number by considering its expansion in simple continued
fractions, but we will not give the proof here.
35
2. Transcendence Criteria
holds for any rational number p/q ∈ Q (with p, q ∈ Z, q > 0 and gcd( p, q) ∼ 1).
We can convince ourselves that the case k = n in the inequality (2.19) gives a
version of Liouville’s Approximation Theorem.
Liouville’s Approximation Theorem has been strengthened over the years by
successively improving the exponent k:
n
Theorem 2.12 (Thue 1909) If n ≥ 3, the inequality (2.19) holds true for k = 2+1
√
Theorem 2.13 (Siegel 1921) If n ≥ 3, the inequality (2.19) holds true for k = 2 n
36
2.2. Liouville’s Approximation Theorem
√
Theorem 2.14 (Dyson 1947) If n ≥ 3, the inequality (2.19) holds true for k = 2n
Theorem 2.15 (Roth 1955) If n ≥ 2, the inequality (2.19) holds true for k = 2
We note that all those theorems are not currently effective, i.e. there is no
bound known on the possible values of p, q given α.
Theorem 2.15 had been conjectured by Siegel in 1921. Even though Thue,
Siegel, and Dyson had successively improved Liouville’s original exponent
n, it was Roth who proved Siegel’s conjectured exponent in 1955, and won
a Fields Medal for this work. Still, Theorem 2.15 is often called the Thue-
Siegel-Roth Theorem. It is the best possible of its kind; the exponent 2 cannot
indeed be decreased, because Roth’s statement would fail on setting ε = 0
due to the Asymptotic Dirichlet’s Theorem.
p 1
α− < 2
q q log(q)1+ε
37
Chapter 3
Transcendence of e
3.1 Preliminaries
We start by introducing an auxiliary function which will happen to be very
useful in the proof of the transcendence of e.
γ : [0, 1] → C
θ 7→ γ(θ ) := θt
Remark 3.2 By Cauchy’s theorem, since the function z 7→ et−z f (z) is holomorphic
on the domain C, the complex line integral I (t; f ) is the same over any piecewise
continuously differentiable path joining 0 and t in C and depends only on the
endpoints 0 and t. That is why we can simply write without ambiguity:
Z t
I (t; f ) = et−z f (z) dz
0
39
3. Transcendence of e
Proof We recall the formula about partial integration: Let v, w be two holo-
morphic functions on a domain G ⊆ C. Let γ : [ a, b] → G be continuous and
piecewise differentiable. Then the following holds:
Z Z
0 γ(b)
v(z)w (z)dz = [v(z)w(z)]γ(a) − v0 (z)w(z)dz
γ γ
By simply applying this formula to the line γ joining 0 and t and to the
following functions:
dv = f 0 (z)dz
v := f (z)
dw := et−z dz w = − et−z
we get:
Z
I (t; f ) = v(z)w0 (z)dz
γ
Z t
= [−et−z f (z)]0t − (−et−z ) f 0 (z)dz
0
= −et−t f (t) + et−0 f (0) + I (t; f 0 )
= et f (0) − f (t) + I (t; f 0 ),
as desired.
By iterating the process in Lemma 3.3, we obtain for any polynom f ∈ R[ X ]
of degree m:
Lemma 3.4 With the same notations as before, and for any polynom f ∈ R[ X ] of
degree m, we have:
m m
I (t; f ) = et ∑ f ( j) (0) − ∑ f ( j) (t)
j =0 j =0
00
Proof We apply Lemma 3.3 to f , f 0 , f , . . . , f (m) to obtain:
40
3.1. Preliminaries
I (t; f ) = et f (0) − f (t) + et f 0 (0) − f 0 (t) + et f 00 (0) − f 00 (t) + . . . + et f (m) (0) − f (m) (t)
as desired.
Now we are interested in finding an upper bound for | I (t; f )| when f ∈ R[ X ]
is a polynomial with real coefficients of degree m:
Lemma 3.5 With the same notations as before, and for any polynom f ∈ R[ X ] of
degree m, we have:
Proof The proof is based on the triangular inequality for integrals. We have
by definition of the auxiliary function:
Z 1 Z 1
| I (t; f )| = | et−θt f (θt)t dθ | ≤ |et−θt | · | f (θt)| · |t| dθ
0 0
But we know that |ez | = eRe(z) and Re(z) ≤ |z| for all z ∈ C, so we obtain for
all θ ∈ [0, 1]:
Now, for the term | f (θt)|, we have for all θ ∈ [0, 1]:
f = f ( X ) = X p −1 ( X − 1 ) p . . . ( X − n ) p
deg( f ) = p − 1 + np = (n + 1) p − 1
41
3. Transcendence of e
We also note that the roots of f are 0, 1, 2, . . . , n. The root 0 is a root of order
p − 1, and the root k for k = 1, 2, . . . , n is a root of order p. Therefore we
certainly have:
f ( j) (k ) = 0 for 0 ≤ j < p (k = 1, 2, . . . , n)
Proof The first case when 0 ≤ j ≤ p − 2 follows from the order p − 1 of the
root 0 of f , as already discussed.
We note that f is of the form: f ( X ) = X p−1 · ( g( X )) p with
g ( X ) = ( X − 1) · . . . · ( X − n ) ∈ Z[ X ]
p −1
p−1
f ( p −1)
(X) = ∑ r
( X p−1 )(r) · (( g( X )) p )( p−1−r)
r =0
( X p −1 ) (r ) 6= 0 only when r = p − 1,
X =0
( X p −1 ) ( p −1) = ( p − 1) !
X =0
42
3.1. Preliminaries
p −1
p−1
f p −1
(X) = ∑ ( X p −1 ) (r ) · (( g( X )) p )( p−1−r)
X =0 r =0 r X =0 X =0
p −2
p−1
= ∑ ( X p −1 ) (r ) · (( g( X )) p )( p−1−r)
r =0 r X =0 X =0
as desired.
Now, for the third case when j ≥ p, we get again by the Leibniz’s formula:
j
j
f ( j) ( X ) = ∑ r (X p−1 )(r) · (( g(X )) p )( j−r)
r =0
Since we have that ( X p−1 )(r) ≡ 0 for r ≥ p, the sum in the formula for the jth
derivative of f with j ≥ p is reduced to the following sum:
p −1
j
f ( j)
(X) = ∑ r
( X p−1 )(r) · (( g( X )) p )( j−r)
r =0
p −1
j
f ( j)
(X) = ∑ r
( X p −1 ) (r ) · (( g( X )) p )( j−r)
X =0 r =0 X =0 X =0
But since the term ( X p−1 )(r) is not zero only for r = p − 1, in which case
X =0
it is equal to ( p − 1)!, we get:
j
f ( j)
(X) = ( X p −1 ) ( p −1) · (( g( X )) p )( j−( p−1))
X =0 p−1 X =0 X =0
j
= ( p − 1)! · (( g( X )) p )( j−( p−1))
p−1 X =0
43
3. Transcendence of e
So for i = j − ( p − 1) > 0 (in the case when j ≥ p), we get for the evaluation
at X = 0:
as desired.
We can also give a formula for the successive derivatives of f at any other
roots k = 1, 2, . . . , n:
44
3.2. The Original Proof of Hermite
The second case, when j ≥ p, is the same as before. Indeed, as in the proof
of Lemma 3.6, we arrive at the conclusion:
j p −1 0 ( j− p) j
( j)
f (k) = ( p − 1)! · p · [( g(k)) · g (k)] = ( p − 1)! · p · l˜
p−1 p−1
g = g ( X ) = a n X n + . . . + a1 X + a0 ∈ Z[ X ]
g ( e ) = a n · e n + . . . + a1 · e + a0 = 0
We can further assume that a0 6= 0, because otherwise, g would not have any
constant term and hence g would have 0 as a root, which is a contradiction,
because any irreducible polynomial in Q[ X ] of degree at least 2 has no
rational root.
We now consider the previous polynomial f ∈ Z[ X ], namely:
f = f ( X ) = X p −1 ( X − 1 ) p . . . ( X − n ) p ,
but only for a large positive prime number p, such that p > n and also such
that p > | a0 |.
Let m denote the degree of f , i.e. m = (n + 1) p − 1. We certainly have, since
p ≥ 2, that m ≥ p.
We consider the following combination involving the auxiliary function I (t; f )
defined in (3.1) for t = 0, 1, 2, . . . , n:
n
J := a0 · I (0; f ) + a1 · I (1; f ) + . . . + an · I (n; f ) = ∑ ak · I (k; f )
k =0
45
3. Transcendence of e
as claimed.
CLAIM 1: J ∈ Z
This is clearly true, using Equation (3.1), since all the ai ’s are integer and
since f ∈ Z[ X ] implies that f ( j) ∈ Z[ X ] (for j = 0, 1, . . . , m), which in turn
implies that the evaluation of f ( j) at the integers k = 0, 1, . . . , n remains an
integer.
CLAIM 2: J 6= 0
We have by taking out the term for k = 0 in Equation (3.1):
m n m
J = − a0 ∑ f ( j ) (0) − ∑ ak ∑ f ( j) ( k )
j =0 k =1 j =0
If J = 0, we would have:
m n m
− a0 ∑ f ( j ) (0) = ∑ ak ∑ f ( j) ( k ) (3.2)
j =0 k =1 j =0
We look at the left-handside of (3.2) and use the formula for the successive
derivatives of f at 0 from Lemma 3.6 (and noting that m ≥ p) to obtain:
" #
m m
− a0 ∑ f ( j) (0) = − a0 · ( p − 1)!(−1)np (n!) p + ∑ f ( j) (0)
j =0 j= p
46
3.2. The Original Proof of Hermite
We claim that this integer composed of two terms is not divisible by p!.
Indeed, the second term
m
− a0 · ∑ f ( j ) (0)
j= p
p −1
! !
n m n m n m
∑ ak ∑ f ( j) ( k ) = ∑ ak ∑ f ( j) ( k ) + ∑ f ( j) ( k ) = ∑ ak 0+ ∑ f ( j) ( k )
k =1 j =0 k =1 j =0 j= p k =1 j= p
Equations (3.3) and (3.4) lead to a contradiction, since both handsides of (3.2)
should be divisible by the same integer, which implies that J 6= 0, as desired.
CLAIM 3: J is divisible by ( p − 1)!
As before, we take out the term for k = 0 in (3.1):
m n m
J = − a0 ∑ f ( j ) (0) − ∑ ak ∑ f ( j) ( k )
j =0 k =1 j =0
47
3. Transcendence of e
| J | ≥ ( p − 1) ! (3.5)
So we have that:
n n
| J| ≤ ∑ |ak || I (k; f )| ≤ ∑ |ak ||k|e|k| supξ ∈C,|ξ |≤|k| | f (ξ )| (3.6)
k =0 k =0
Moreover, we have:
| f (ξ )| = |ξ | p−1 · |ξ − 1| p . . . |ξ − n| p
|ξ − i | ≤ |ξ | + i ≤ n + i ≤ 2n,
48
3.2. The Original Proof of Hermite
m = p − 1 + np ≤ p + np ≤ np + np = 2np
So we have:
supξ ∈C,|ξ |≤n | f (ξ )| ≤ (2n)2np = ((2n)2n ) p (3.8)
Hence, we have, using (3.6), (3.7) and (3.8):
n
| J | ≤ A · n · en · ((2n)2n ) p · ∑ 1 = A · en · n · ((2n)2n ) p · (n + 1)
k =0
c2 c2 ci
e c2 = 1 + + 2 +...+ 2 +...
1! 2! i!
is a convergent series. Hence, for its general term, we have:
c2i
→ 0 as i → ∞
i!
In particular, since c1 > 0, if p is large enough, then ( p − 1) is large enough
to have:
p −1
c2 1
<
( p − 1) ! c1
So we get
p −1
c2
c1 · <1 (3.10)
( p − 1) !
Hence, for p prime large enough such that p > n and p > | a0 | and such that
p −1
c
c1 · ( p2−1)! < 1 (this is possible since there are infinitely many primes), we get,
using (3.5) and (3.9):
p −1
p −1 p −1 c
( p − 1) ! ≤ | J | ≤ c1 · c2 =⇒ ( p − 1)! ≤ c1 · c2 =⇒ c1 · 2 ≥1
( p − 1) !
which is a contradiction to (3.10). Hence, e is not an algebraic number, hence
e is a transcendental number.
49