Number Theory and Combinatorics Compress
Number Theory and Combinatorics Compress
and
Combinatorics
B SURY
Prof. Sury has contributed significantly to research in the areas of linear al-
gebraic groups over global and local fields, Diophantine equations, division
algebras, central extensions of p-adic groups, applications of density theo-
rems in number theory, K-theory of Chevalley groups, combinatorial num-
ber theory, and generation of matrix groups over rings. The book, which
will be available in digital format, and will be housed as always on the
Academy website, will be valuable to both students and experts as a useful
handbook on Number Theory and Combinatorics.
Amitabh Joshi
Editor of Publications
Indian Academy of Sciences
August 2017
iii
About the Author
Dipendra Prasad
TIFR, Mumbai
v
Contents
vii
Preface
ix
as not easy to understand. In almost all cases, after a meeting or discussion
with some students, I felt compelled to write on a certain topic. Inciden-
tally, some of the summer projects that students worked on have appeared
in Resonance as well and, many of them were rewritten by me – showing
my desire to communicate in a particular way. I personally know several
colleagues who are much better at writing such expositions but, not many
of them seem to find the time to write at this level. I wish they would.
Finally, I am indebted to Professor Ram Ramaswamy for suggesting the
publication of this volume, for convincing me that it could be useful, and
for pushing me to finish this process.
B Sury
July 2017
x
Cyclotomy and Cyclotomic Polynomials:
The Story of how Gauss Narrowly
Missed Becoming a Philologist
1
Cyclotomy and Cyclotomic Polynomials
2
Cyclotomy and Cyclotomic Polynomials
l
B
P l
B
A
3
Cyclotomy and Cyclotomic Polynomials
Even when the roots of x2 − ax + b = 0 are not real, they can √be con-
2
structed easily. In this case, we need to construct the points ( a2 , ± 4b−a
2 ).
But, as we
(0,1)
(0,0) x-axis
( a–√a2 –4b ,0) ( a+√a2 –4b ,0)
2 2
(a,b)
y-axis
0 = ζ + ζ 2 + ζ 3 + ζ 4 + ζ 5.
(ζ 2 + ζ 3 )(ζ + ζ 4 ) = ζ + ζ 2 + ζ 3 + ζ 4 = −1.
(ζ 2 + ζ 3 ) + (ζ + ζ 4 ) = −1.
4
Cyclotomy and Cyclotomic Polynomials
This means that ζ 2 +ζ 3 and ζ +ζ 4 are the two roots of the quadratic poly-
nomial T 2 +√
T − 1 = 0. Thus, ζ + ζ (being positive, equal to 2Cos(2π/5))
−1+ 5
equals 2 . Multiplying this equality by ζ and using ζ 5 = 1, one gets a
quadratic equation for ζ!
This is the algebraic reasoning behind the construction. Following it, we
can geometrically make the construction also with the aid of the dictionary
between algebra and geometry that we have established above.
(– 12 ,√5
2 )
(–1,0) (1,0)
(– 12 , – √5
2 )
Let us now turn to the construction of the 17-gon. There are many ways
of doing it – [4], [5] contain explicit geometric algorithms; Gauss’s own
construction appears in [6], Art.365 – and all of them succeed essentially
because 17−1 is a power of 2! This is the reason why the degree 16 equation
x17 −1 16 15
x−1 = x +x +· · ·+x+1 = 0 reduces to a chain of quadratic equations.
It would be ideal to use the language of Galois theory (see Resonance,
Vol. 4, No. 10, 1999) to discuss the constructibility or non-constructibility of
a regular polygon. However, we will keep the discussion elementary and will
only make a few remarks for the reader familiar with basic Galois theory
so that she/he can grasp the conceptual reason behind various explicit
expressions, the appearance of which will seem magical without the added
understanding4 provided by Galois theory.
In the light of our dictionary, we describe a construction as follows:
Denote by ζ, the 17-th root of unity e2iπ/17 . Then, ζ 17 = 1 gives
ζ 16 + ζ 15 + · · · + ζ + 1 = 0.
That is,
(ζ + ζ −1 ) + (ζ + ζ −2 ) + · · · + (ζ 8 + ζ −8 ) = −1.
4 Perhaps
an outstanding feature of mathematics is that knowing the conceptual reason behind
a phenomenon is often much more important than a proof of the phenomenon itself.
5
Cyclotomy and Cyclotomic Polynomials
6
Cyclotomy and Cyclotomic Polynomials
7
Cyclotomy and Cyclotomic Polynomials
(–1,0) (1,0)
3. Cyclotomic Polynomials
We introduced for any positive integer n, the cyclotomic polynomial
Φn (X) as the unique monic integer polynomial of least degree having ζ =
e2iπ/n as a root. What does Φn (X) look like? Obviously Φ1 (X) = X − 1
8
Cyclotomy and Cyclotomic Polynomials
9
Cyclotomy and Cyclotomic Polynomials
question is whether there are infinitely many of each type? The answer is
‘yes’ by a deep theorem due to Dirichlet – infinitely many primes occur in
any arithmetic progression {a + nd; n ≥ 1} with a, d coprime.
If d is a positive integer, then for the arithmetic progression {nd + 1;
n ≥ 1}, one can use cyclotomic polynomials to prove this! This is not
surprising because we have already noted in the last section that cyclotomic
polynomials are related to the way prime numbers are distributed. Let us
prove this now.
Suppose p1 , p2 , · · · , pr are prime numbers in this arithmetic progression.
We will use cyclotomic polynomials to produce another prime p in this
progression different from the above pi ’s. This would imply that there are
infinitely many primes in such a progression. We will use the simple obser-
vation that a polynomial p(X) with integer coefficients has the property
that p(m) − p(n) is an integer multiple of m − n.
Consider the number N = dp1 p2 · · · pr . Then, for any integer n, the two
values Φd (nN ) and Φd (0) differ by a multiple of N . But, Φd (0) is an integer
which is also a root of unity and must, therefore, be ±1. Moreover, as
n → ∞, the values Φd (nN ) → ∞ as well since Φd is a nonconstant monic
polynomial. In other words, n > 0, the integer Φd (nN ) has a prime factor
p. As Φd (nN ) is ±1 modulo any of the p1 , p2 , · · · , pr and modulo d, the
prime p is different from any of the pi ’s and does not divide d. One might
wonder which primes divide some value Φd (a) of a cyclotomic polynomial.
The answer is that these are precisely the primes occurring in the arithmetic
progression {nd + 1; n > 0}. To show this, we use the idea that the nonzero
integers modulo p form a group of order p − 1 under the operation of
multiplication modulo p. So, it is enough to prove that if p divides Φd (a)
for some integer a, then a has order d in this group (for, then Lagrange’s
theorem of finite group theory tells us that d divides the order p − 1 of
the group, which is just re-stating that p is in the arithmetic progression
{nd+1; n > 0}. Let us prove this now. Since X d −1 = l|d Φl (X), it follows
Q
that p which divides Φd (a) has to divide ad − 1 also. If d were not the order
divide d with k < d and p divides ak − 1. Once again, the relation
of a, let k Q
k
a − 1 = l|k Φl (a) shows that p divides Φl (a) for some positive integer l
dividing k. Therefore, p divides both Φd (a + p) and Φl (a + p). Now,
Y
(a + p)d − 1 = Φm (a + p) = Φd (a + p)Φl (a + p) (other terms).
m|d
10
Cyclotomy and Cyclotomic Polynomials
occurs in the arithmetic progression {1+nd; n > 0} and thereby, proves the
infinitude of the primes in this progression. Interestingly, Euclid’s classical
proof of the infinitude of prime numbers is the special case of the above
proof where we can use d = 2.
11
Cyclotomy and Cyclotomic Polynomials
is called a Ramanujan sum. In other words, it is simply the sum of the k-th
powers of the primitive n-th roots of unity – ‘primitive’ here means that the
number is not an m-th root of unity for any m < n. Note that the primitive
n-th roots of unity are the numbers e2ikrπ/n for all those r ≤ n which are
relatively prime to n.
The first remarkable property cn (k) have is that they are integers. Ra-
manujan showed that several arithmetic functions (that is, functions de-
fined from the set of positive integers to the set of complex numbers) have
‘Fourier-like’ of expansions in terms of the sums; hence, nowadays these ex-
pansions are known as Ramanujan expansions. They often yield very pretty
elementary number-theoretic identities. Recently, the theory of group rep-
resentations of the permutation groups (specifically, the so-called super-
character theory has been used to re-prove old identities in a quick way
and also, to discover new identities.
It is convenient to write
Then, the set of all n-th roots of unity {e2ikπ/n : 0 ≤ k < n} is a union
of the disjoint sets ∆d as d varies over the divisors of n. This is because an
n-th root of unity is a primitive d-th root of unity for a unique divisor d of
n. It is also convenient to introduce the ‘characteristic’ function δk|n which
has the value 1 when k divides n and the value 0 otherwise. Before stating
some properties of the ck (n)’s, let us recall two arithmetic functions which
are ubiquitous in situations where elementary number-theoretic counting is
involved. The first one is Euler’s totient function
then
X
g(n) = f (d)µ(n/d).
d|n
12
Cyclotomy and Cyclotomic Polynomials
With these notations, here are some elementary properties of the Ra-
manujan sums.
(i) cn (k) = cn (−k) = cn (n − k).
(ii) cn (0) = φ(n) and cn (1) = µ(n).
(iii) cn (ks) = cn (k) if (s, n) = 1; in particular, cn (s) = µ(n) if (s, n) = 1.
(iv) cn (k) = cn (k 0 ) if (k, n) = (k 0 , n); in particular, cn (k) ≡ cn (k 0 ) mod n
if k ≡ k 0 mod n.
P n−1
(v) Pk=0 cn (k) = 0. P P
(vi) d|n cd (k) = δn|k n and cn (k) = d|n dµ(n/d)δd|k = d|(n,k) dµ(n/d);
in particular, for prime powers pr , we have cpr (k) = pr − pr−1 if pr |k;
= −pr−1 if pr−1 ||k; and = 0 otherwise.
(vii) cmn (k) = cm (k)cn (k) if (m, n) = 1.
P n
(viii) k=1 cm (k)cn (k) = δmn nφ(n).
The property (vi) shows that these sums actually have integer values.
The proof of (i) follows already from the definition and, so do the first
parts of (ii) and (iii). The second parts of (ii), (iii) as well as the assertions
(iv) and (vii) will follow from (vi). We shall prove (v) and (vi).
For (v), we have
n−1
X n−1
X X X n−1
X
k
cn (k) = ζ = ζk = 0
k=0 k=0 ζ∈∆n ζ∈∆n k=0
for each ζ ∈ ∆n .
For proving (vi), we note that the second statement follows from the first
by the Möbius inversion formula. Let us prove the first one now. We have
X X X n−1
X
k
cd (k) = ζ = e2imkπ/n
d|n d|n ζ∈∆d m=0
13
Cyclotomy and Cyclotomic Polynomials
The right hand side was studied by R D Von Sterneck in 1902 and is
known by his name. The equality above itself was known before Ramanujan
and is due to J C Kluyver in 1906.
Φ0n (x) X
x(xn − 1) = dµ(n/d)(xd + x2d + · · · + xn ).
Φn (x)
d|n
7 Note
that even computationally the defining sum for cn (k) requires approximately n operations
whereas the other sum requires roughly log(n) operations.
14
Cyclotomy and Cyclotomic Polynomials
13 + 33 + · · · + n3 = (1 + 2 + · · · + n)2 .
(13 + 23 + 33 + · · · + n3 )k = (1 + 2 + 3 + · · · + n)2k .
Are there other such identities? It turns out that there are no others. We
shall prove this now using cyclotomic polynomials. To be more precise, let
us set up the notation
pr (n) = 1r + 2r + · · · + nr .
for any natural numbers n, r. Let us look for natural numbers r1 < r2 <
· · · < rk and s1 < s2 < · · · < sl different from the ri ’s and, also some natural
numbers a1 , a2 , · · · , ak , b1 , b2 , · · · , bl such that, for any natural number n,
one has identities
pr1 (n)a1 pr2 (n)a2 · · · prk (n)ak = ps1 (n)b1 ps2 (n)b2 · · · psl (n)bl .
p3 (n)a = p1 (n)2a ,
that is,
(13 + 23 + 33 + · · · + n3 )a = (1 + 2 + 3 + · · · + n)2a
15
Cyclotomy and Cyclotomic Polynomials
which we have already seen. This is an easy check. Now, let us suppose
sl > 3. Below, we will prove the nice fact that any number of the form
1 + 2b with b > 3 always has a prime factor which is not a factor of any
1 + 2c for any c < b. This beautiful observation was first made by A S
Bang 120 years ago. This observation shows immediately that an equality
of the form (A) cannot hold good because the prime factor p of the largest
1 + 2sl cannot divide any term on the left hand side. Seeing why Bang’s
observation is valid requires some discussion about cyclotomic polynomials
which we proceed to do now.
Generally, if one has an infinite sequence of natural numbers u1 < u2 <
u3 < · · · such that for every n, there exists a prime factor of un which does
not divide um for every m < n usually called a primitive prime divisor of
un . From any such sequence admitting primitive prime divisors, we have a
proof of infinitude of primes because we find at least one new prime divisor
at each step as we move along the sequence of un ’s.
We show now that the sequence {2n + 1}n>3 has primitive divisors.
An advantage of knowing that the cyclotomic polynomials Φn (x) have in-
teger coefficients is the following. For any integer a and any natural number
n, one has
Y
an − 1 = Φd (a)
d|n
16
Cyclotomy and Cyclotomic Polynomials
17
Cyclotomy and Cyclotomic Polynomials
i−1
where b = 2p and ζr are the φ(d) primitive d-th roots of unity (the roots
of Φd (x)).
As bp − 1 ≥ bp−2 (b2 − 1), the right side above is > b(p−2)φ(d) (b − 1)φ(d) .
As b ≥ 2, this last expression is at least 2p−2 . Therefore, we have
p = Φ2n (2) > 2p−2 ,
which is possible only if p = 3. In that case we must also have 2n = 6
which we rule out. In other words, when n > 3, then there does exist
a prime divisor p of Φ2n (2) which does not divide n; the above discussion
then shows that n is the smallest natural number for which p divides 2n +1.
This finishes the whole argument.
··· + X n−1 has as roots all the n-th roots of 1 excepting 1 itself, we have
n−1
Y
(1 − e2idπ/n ) = n,
d=1
18
Cyclotomy and Cyclotomic Polynomials
First, let us look at the case when n = pk for some prime p. Then,
Qpk −1 k
Y
2idπ/pk |1 − e2idπ/p | pk
|1 − e | = Q d=1 k = = p.
dp<pk |1 − e
2idpπ/p | pk−1
(d,pk )=1,d<pk
If p is a prime dividing
Qn−1 n, suppose pk is the highest power of p dividing n.
Then, the product d=1 (1 − e2idπ/n ) contains the products of terms corre-
Qpk −1 k
sponding to d running through multiples of n/pk ; that is, d=1 (1−e2idπ/p )
(which is pk ). We observe that factors occurring for a different prime q
dividing n are disjoint from those occurring corresponding to p. There-
fore,
Q thek factors corresponding to the various primes dividing n contribute
pk ||n p = n.
On removing these factors corresponding to each prime divisor of n, we
will get d∈D (1 − e2idπ/n ) = 1, where D consists of those d for which
Q
e2idπ/n does not have prime power order. Thus, if d ∈ D, then 1 − e2idπ/n
is a unit since n is not a prime power. Therefore, 1 − e2iπ/n is a unit in the
cyclotomic field Q(e2iπ/n ). From Galois theory, we have that the product
2idπ/n ) is the norm of 1 − e2iπ/n from Q(e2iπ/n ) to Q. As this
Q
(d,n)=1 (1 − e
element is a unit, this product is ±1. Hence we get (d,n)=1 |1−e2idπ/n | = 1
Q
which proves our assertion in the case when n is not a prime power. The
proof is complete.
In the above proof, the second part can also be deduced from the first
part of the proof inQa different fashion as follows.
Writing P (n) = n−1 l d
Q
l=1 (1 − ζ ) and Q(n) = (d,n)=1 (1 − ζ ), where ζ =
e2iπ/n , we can see that
Y
P (n) = Q(r).
r|n
can be identified with the so-called von Mangoldt function Λ(n) which
is defined to have the value log(p) if n is a power of p and 0 otherwise.
Using this identification, exponentiation gives also the value asserted in the
proposition; viz., Q(n) = p or 1 according as to whether n is a power of a
prime p or not.
19
Cyclotomy and Cyclotomic Polynomials
vp (n)
P Q
To see why Λ(n) = d|n µ(n/d) log(d), we write n = p|n p and note
that
X
log(n) = vp (n) log(p)
p|n
P
But, the right hand side is clearly d|n Λ(d). Hence, Möbius inversion
yields
X
Λ(n) = log(d)µ(n/d).
d|n
we may write
Xn − 1 Y
Ψ(X) := = (X d − 1)−µ(n/d).
Φn (X)
d|n,d<n
20
Cyclotomy and Cyclotomic Polynomials
n/p
− 1)−1 as units have norm ±1.
Q
So, its norm is ± p|n NK/Q (ζn
n/p
As ζn is a primitive p-th root of unity, it is in the subfield Q(ζp ) gen-
erated by a primitive p-th root of unity, and we have
Thus, we get
nφ(n)
d(K) = ± Q φ(n)/(p−1)
.
p|n p
21
Cyclotomy and Cyclotomic Polynomials
We recall:
For a positive integer n > 2, if disc(Φn ) is a perfect square, then Φn is
reducible modulo every prime.
This proof is a standard application of Galois theory. Indeed, it is well-
known that if the discriminant of a Galois extension is a square, its Galois
group would be contained in the subgroup of even permutations ([1], Lemma
12.3). So, if Φn were irreducible modulo some prime p, then the reduction
of Φn mod p generates over Fp a Galois extension of degree φ(n); the Galois
group would contain a φ(n)-cycle which is an odd permutation since φ(n)
is even for n
We prove:
For n > 2, the polynomial Φn is reducible modulo every prime if, and
only if, disc(Φn ) is a perfect square. If disc(Φn ) is not a perfect square –
which happens if, and only if, n = 4, pk or 2pk – then there are infinitely
many primes p such that Φn is irreducible modulo p.
Let us prove this now.
We have already seen that if disc(Φn ) is a perfect square in Z, then
Φn is reducible modulo every prime. Conversely, suppose disc(Φn ) is not a
φ(n)
perfect square. Then, looking at the expression (−1)φ(n)/2 Q pnφ(n)/(p−1) for
p|n
the discriminant, we shall deduce that n = 4, pk or 2pk for some odd prime.
Indeed, write
n = pα1 1 pα2 2 · · · pαr r .
Firstly, if n is odd and r > 1, clearly,
pαi −1 (pi − 1)
Qr
φ(n)
= i=1 i ,
2 2
is even and the power of pi dividing the discriminant is
Y r Y
αi (pi − 1) − 1 pαk k −1 (pj − 1) ,
k=1 j6=i
which is even.
Thus, if n > 2 is odd, then the discriminant is a perfect square unless
n = pk .
If n = 2pα1 1 · · · pαr r for some odd primes, Φn = Φn/2 and the discriminant
is a perfect square excepting the case r = 1; i.e., n = 2pk .
Now, if n = 2α pα1 1 · · · pαr r with either α > 2 or α = 2 and r ≥ 1, then
again the powers of 2 and each pi dividing the discriminant are all even.
Thus, the exceptional case is n = 4.
Therefore, we have deduced that the expression for discriminant is a
perfect square excepting the cases n = 4, pk and 2pk for an odd prime.
22
Cyclotomy and Cyclotomic Polynomials
These exceptional cases are when the Galois group of the cyclotomic field
is cyclic.
The Galois group of Φn over Q is a cyclic group of order φ(n) and con-
tains a φ(n)-cycle. By the Frobenius density theorem discussed in a later
chapter, there are infinitely many prime numbers l such that the decompo-
sition group at l is cyclic of order φ(n) which means that Φn modulo l is
irreducible and generates the extension of degree φ(n) over Fl . This proves
the proposition.
References
[1] P Morandi, Field and Galois theory, Graduate texts in Mathematics
167, Springer-Verlag, 1996.
[2] M Ram Murty and J Esmonde, Problems in Algebraic Number Theory,
Graduate Texts in Mathematics 190, Springer-Verlag, New York, 2005.
[3] M Artin, Algebra, Prentice Hall, 1991.
[4] D Suryaramana, Resonance, Vol. 2, No. 6, 1997.
[5] Ian Stewart, Gauss, Scientific American, 1977.
[6] C F Gauss, Disquisitiones Arithmeticae, English Edition, Springer,
1985.
[7] J Gray, English translation and commentary on Gauss’s mathematical
diary, Expo. Math., Vol. 2, 1984.
[8] M Rosen, Amer. Math. Monthly, 1981.
[9] B Sury, Ramanujan’s Awesome Sums,Mathematics Newsletter,
Vol. 24, no. 2, pp. 31–36, September 2013.
[10] S Ramanujan, On certain trigonometrical sums and their applications
in the theory of numbers, Trans. Cambridge Philos. Soc., Vol. 22,
No. 13, 259-276, 1918.
23
Polynomials with Integer Values
25
Polynomials with Integer Values
COROLLARY 1.2
If a polynomial P takes integers to integers and has degree n, then n!P (X)
∈ Z[X].
Lemma 1.3. A nonconstant integral polynomial P (X) cannot take only
prime values.
Proof. If all values are composite, then there is nothing to prove. So,
assume that P (a) = p for some integer a and prime p. Now, as P is non-
constant, lim |P (a + np)| = ∞. So, for big enough n, |P (a + np)| > p. But
n→∞
P (a + np) ≡ P (a) ≡ 0 mod p, which shows P (a + np) is composite.
Remark 1.4. Infinitely many primes can occur as integral values of a
polynomial. For example, if (a, b) = 1, then the well-known (but deep)
Dirichlet’s theorem on primes in progression shows that the polynomial
aX +b takes infinitely many prime values. In general, it may be very difficult
to decide whether a given polynomial takes infinitely many prime values.
For instance, it is not known if X 2 + 1 represents infinitely many primes.
In fact, there is no known polynomial of degree ≥ 2 which takes infinitely
many prime values.
Lemma 1.5. If P is a nonconstant polynomial that takes integers to inte-
gers, the number of prime divisors of its value set {P (m)}m∈Z , is infinite
i.e. not all terms of the sequence P (0), P (1), · · · can be built from finitely
many primes.
Proof. It is clear from the note above that it is enough to prove this for
n
ai X i
P
P (X) ∈ Z[X], which we will henceforth assume. Now, P (X) =
i=0
where n ≥ 1. If a0 = 0, then clearly P (p) ≡ 0 mod p for any prime p. If
a0 6= 0, let us consider for any integer t, the polynomial
n n
( )
X X
P (a0 tX) = ai (a0 tX)i = a0 1 + ai a0i−1 ti X i = a0 Q(X).
i=0 i=1
There exists some prime number p such that Q(m) ≡ 0 mod p for some
m and some prime p, because Q can take the values 0,1,−1 only at finitely
many points. Since Q(m) ≡ 1 mod t, we have (p, t) = 1. Then P (a0 tm) ≡
0 mod p. Since t was arbitrary the set of p arising in this manner is infinite.
Remark 1.6. (a) Note that it may be possible to construct infinitely many
terms of the sequence {P (m)}m∈Z using only a finite number of primes.
For example take (a, d) = 1, a ≥ d ≥ 1. Since, by Euler’s theorem, aϕ(d) ≡
ϕ(d)n
1 mod d, the numbers a(a d −1) ∈ Z ∀ n. For the polynomial P (X) =
dX + a, the infinitely many values P ( ad (aϕ(d)n − 1)) = aϕ(d)n+1 have only
prime factors coming from primes dividing a.
26
Polynomials with Integer Values
(b) In order that the values of an integral polynomial P (X) be prime for
infinitely many integers, P (X) must be irreducible over Z and of content 1.
By content, we mean the greatest common divisor of the coefficients. In
general, it is difficult to decide whether a given integral polynomial is irre-
ducible or not. We note that the irreducibility of P (X) and the condition
that it have content 1, are not sufficient to ensure that P (X) takes infinitely
many prime values. For instance, the polynomial X n + 105X + 12 is irre-
ducible, by Eisenstein’s criterion (see Box 1). But, it cannot take any prime
value because it takes only even values and it does not take either of the
values ±2 since both X n + 105X + 10 and X n + 105X + 14 are irreducible,
again by Eisenstein’s criterion.
Lemma 1.7. Let a1 , · · · , an be distinct integers.
Then P (X) = (X − a1 ) · · · (X − an ) − 1 is irreducible.
Proof. Suppose, if possible, P (X) = f (X)g(X) with deg .f, deg .g < n.
Evidently, f (ai ) = −g(ai ) = ±1 ∀1 ≤ i ≤ n. Now, f (X) + g(X) be-
ing a polynomial of degree < n which vanishes at the n distinct integers
a1 , · · · , an must be identically zero. This gives P (X) = −f (X)2 but this is
impossible as can be seen by comparing the coefficients of X n .
Exercise 1.8. Let n be odd and a1 , · · · , an be distinct integers. Prove that
(X − a1 ) · · · (X − an ) + 1 is irreducible.
Let us consider the following situation. Suppose p = an · · · a0 is a prime
number expressed in the usual decimal system i.e. p = a0 + 10a1 + 100a2
+ · · · + 10n an , 0 ≤ ai ≤ 9. Then, is the polynomial a0 + a1 X + · · · + an X n
irreducible? For example 1289 is a prime the following result due to A Cohn
and x3 + 2x2 + 8x + 9 is irreducible. This is, in fact, true more generally
and, we have:
Lemma 1.9. Let P (X) ∈ Z[X] and assume that there exists an integer n
such that
(i) the zeros of P lie in the half plane Re(z) < n − 21 .
(ii) P (n − 1) 6= 0
(iii) P (n) is a prime number.
Then P (X) is irreducible.
Proof. Suppose, if possible P (X) = f (X)g(X) over Z with f, g hav-
ing positive degrees. All the zeros of f (X) also lie in Re(z) < n − 12 .
Writing f as a product of its irreducible factors over R, we can observe
that f (x + n − 1/2) has ALL coefficients non-zero and of the same sign.
Thus, the coefficients of f (−x + n − 1/2) have alternate signs. Therefore,
|f (n − 21 − t)| < |f (n − 12 + t)|∀t > 0. Since f (n − 1) 6= 0 and f (n − 1) is
integral, we have |f (n − 1)| ≥ 1. Thus |f (n)| > |f (n − 1)| ≥ 1. A similar
27
Polynomials with Integer Values
thing holding for g(X), we get that P (n) has proper divisors f (n), g(n)
which contradicts our hypothesis.
Remark: Michael Filaseta and collaborators have generalized
this vastly. They show that there exists an integer polynomial f of
degree 129 explicitly written down whose largest coefficient is
49598666989151226098104244512919 such that f (10) is prime but f has
the factor x2 − 20x + 101. Further, every integer polynomial g of any degree
whose coefficients are non-negative and strictly less than the above number
must be irreducible if g(10) is prime!
28
Polynomials with Integer Values
The second and third equalities above show that P (X) is reducible mod-
ulo any pn and any q n . Also since p ≡ 1 mod 8, p is a quadratic residue
modulo any 2n and the second equality above again shows that P (X) is
the difference of two squares modulo 2n , and hence reducible mod 2n .
If ` is a prime 6= 2, p, q, let us show now that P (X) is reducible modulo
ln for any n.
At least one of ( p` ), ( q` ) and ( pq
` ) is 1 because, by the product formula for
p
Legendre symbols, ( ` ) · ( ` ) · ( ` ) = 1. According as ( p` ), ( q` ) or ( pq
q pq
` ) = 1, the
second, third or fourth equality shows that P (X) is reducible mod `n for
any n.
We end this section with a result of Schur whose proof is surprising and
elegant as well. This is:
Schur’s Theorem 2.4. For any n, the truncated exponential polynomial
2 n
En (X) = n!(1 + X + X2! + · · · + Xn! ) is irreducible over Z.
Just for this proof, we need some nontrivial number theoretic facts. A
reader unfamiliar with these notions but one who is prepared to accept at
face value a couple of results can still appreciate the beauty of Schur’s proof.
Here is where we have to take recourse to some very basic facts about prime
decomposition in algebraic number fields. Start with any (complex) root α
of f and look at the field K = Q(α) of all those complex numbers which
can be written as polynomials in α with coefficients from Q. The basic
fact that we will be using (without proof) is that any nonzero ideal in ‘the
ring of integers of K’ (i.e., the subring OK of K made up of those elements
29
Polynomials with Integer Values
30
Polynomials with Integer Values
31
Polynomials with Integer Values
Proof. If P (X) is not an exact k-th power, then one can write P (X)
= f (X)k g(X) for polynomials f, g so that g(X) has a zero whose multiplic-
ity is < k. Once again, we can choose n and a prime p such that g(n) ≡ 0
mod p, 6≡ 0 mod pk . This contradicts the fact that P (n) is a k-th power.
Remark: The above results and much more general properties of polyno-
mials are consequences of the so-called Hilbert irreducibility criterion which
implies: if f (X, Y ) is an irreducible polynomial with rational coefficients,
then there exist infinitely many rational values a of x such that the polyno-
mials f (a, Y ) are irreducible in Q[Y ]. One application of the above theorem
is:
Given two non-constant polynomials f, g with rational coefficients such
that f (Q) is contained in g(Q), there exist a polynomial h with rational
coefficients so that f (X) = g(h(X)).
4. Cyclotomic Polynomials
These were already referred to in the earlier chapter. It was also shown
there that one could use these polynomials to prove the existence of in-
finitely many primes congruent to 1 modulo n for any n. For a natural
number d, recall that the cyclotomic polynomial Φd (X) is the irreducible,
monic polynomial whose roots are the primitive d-th roots of unity i.e.
Φd (X) = a≤d:(a,d)=1 (X − e2πa/d ). Note that Φ1 (X) = X − 1 and that
Q
32
Polynomials with Integer Values
PROPOSITION 4.3.
Every integer occurs as a coefficient of some cyclotomic polynomial.
Proof. First, we claim that for any integer t > 2, there are primes p1 < p2
< · · · < pt such that p1 + p2 > pt . Suppose this is not true. Then, for some
t > 2, every set of t primes p1 < · · · < pt satisfies p1 + p2 ≤ pt . So, 2p1 < pt .
Therefore, the number of primes between 2k and 2k+1 for any k is less than
t. So, π(2k ) < kt. This contradicts the prime-number theorem as noted
above. Hence, it is indeed true that for any integer t > 2, there are primes
p1 < p2 < · · · < pt such that p1 + p2 > pt .
Now, let us fix any odd t > 2. We shall demonstrate that both −t + 1 and
−t + 2 occur as coefficients. This will prove that all negative integers occur
as coefficients. Then, using the fact that for an odd m > 1, Φ2m (X) =
Φm (−X), we can conclude that all integers are coefficients.
Consider now primes p1 < p2 < · · · < pt such that p1 + p2 > pt . Write
pt = p for simplicity. Let n = p1 · · · pt and let us write Φn (X) modulo X p+1 .
Since X n − 1 = d/n Φd (X), and since p1 + p2 > pt , we have
Q
t
Y 1 − X pi
Φn (X) ≡ ≡ (1 + · · · + X p )(1 − X p1 ) · · · (1 − X pt )
1−X
i=1
≡ (1 + · · · + X p )(1 − X p1 − · · · − X pt ) mod X p+1 .
Therefore, the coefficients of X p and X p−2 are 1−t and 2−t respectively.
This completes the proof. Note that in the proof, we have used the fact that
if P (X) = (1−X r )Q(X) for a polynomial Q(X), then Q(X) = P (X)(1+X r
+ X 2r + · · · + · · · ) modulo any X k .
33
Polynomials with Integer Values
Exercise 4.4.
(a) Let A = (aij ) be a matrix in GL(n, Z) i.e., both −1 have integer
Pn A and Aj
entries. Consider the polynomials pi (X) = j=0 aij X for 0 ≤ i ≤ n.
Prove that any integral polynomial of degree at most n is an integral
linear combination of the pi (X). In particular, if a0 , . . . , an ∈ Q are
distinct, show that any rational polynomial of degree at most n is of
n
λi (X + ai )n for some λi ∈ Q.
P
the form
i=0
[n
2
]
+ . . . + Xn (−1)i n−i
P i n−2i . Conclude
(b) Prove that 1 + X = i X (1 + X)
i=0
P n−i 1+ρ+···+ρn
that i = (1+ρ)n where ρ is either root of X 2 + 3X + 1 = 0.
i≥0
(−1)i n−i
P
Further, compute i .
i≥0
n−i
P
Remark 4.5. It is easily seen by induction that i is just the
i≥0
(n+1)-th Fibonacci number Fn+1 . Thus, exercise (b) provides an expression
for Fn+1 . This expression makes it easy to prove the following identities:
(a) Fn + Fn+1 = Fn+2 .
(b) F
Pn+1 Fn−1 = Fn2 + (−1)n .
n z
(c) Fn z = 1−z−z 2.
(d) 2 2
Fn + Fn+1 = F2n+1 .
n−i
P
Notice that only (a) seems obvious from the expression Fn+1 = i .
i≥0
As we remarked earlier, even for a polynomial of degree 2 (like X 2 + 1)
it is unknown whether it takes infinitely many prime values. A general
conjecture in this context is:
Conjecture 4.6. (Bouniakowsky, Schinzel and Sierpinski.) A nonconstant
irreducible integral polynomial whose set of values has no nontrivial
common factor, always takes on a prime value.
It is appropriate to recall here that the polynomial X 2 + X + 41 takes
prime values at X = −40, −39, . . . , 0, 1, . . . , 39. We end with an open ques-
tion which is typical of number-theoretic questions – a statement which
can be understood by the proverbial layman but an answer which proves
elusive to this day to professional mathematicians.
For any irreducible, monic,
Q integral polynomial P (X), define its Mahler
measure to be M (P ) = i Max(|αi |, 1) where the product is over the roots
of P . The following is an easy exercise.
Exercise 4.7. M (P ) = 1 if, and only if, P is cyclotomic.
D H Lehmer posed the following question:
34
Polynomials with Integer Values
Does there exist a constant C > 0 such that M (P ) > 1 + C for all
noncyclotomic (irreducible) polynomials P ?
This is expected to have an affirmative answer and, indeed, Lehmer’s
calculations indicate that the smallest possible value of M (P ) 6= 1 is
1.176280821...., which occurs for the polynomial
P (X) = X 10 + X 9 − X 7 − X 6 − X 5 − X 4 − X 3 + X + 1.
References
[1] Polya and Szego, Problems in Analysis, I & II, Springer-Verlag, 1945.
35
Polynomials with Integer Values
a0 = b0 c0 , a1 = a0 b1 + b0 a1 , · · · , an = br cs , r + s = n.
36
How Far Apart are Primes?
Bertrand’s Postulate
1 → 2, 2 → 3, 3 → 5, 4 → 5, 5 → 7, 6 → 7, 7 → 11 etc.
Thus, it seems that we need to go ‘at most twice the distance’ i.e., we
seem to be able to find a prime between n and 2n for the first few values of
n. But there is absolutely no pattern here. In fact, although we have seen
above that there are arbitrarily large gaps between primes, it is nevertheless
true that ‘there is regularity in the distribution of primes’. It is this fascinat-
ing clash of tendency which seems to make primes at once interesting and
intriguing. It turns out indeed to be true that: there is always a prime be-
tween n and 2n. This statement, known as Bertrand’s postulate, was stated
by Bertrand (1822–1900) in 1843 and proved later by Chebychev in 1852.
Actually, Chebychev proves a much stronger statement which was further
generalised to yield a fundamental fact about the prime numbers known
as the prime number theorem. Interestingly, Bertrand’s motivation was to
group theory and not really number theory; he made many contributions
to differential geometry and probability theory as well.
Proving the prime number theorem is beyond the scope of this article
but stating it certainly lies within it. For a positive real number x, let us
denote by π(x), the number of primes which do not exceed x. The prime
number theorem states that the ratio π(x)logx x approaches the limit 1 as x
π(x)logx
grows indefinitely large i.e., limx→∞ x = 1.
x
One usually writes π(x) ∼ logx to describe such an asymptotic result.
What Chebychev proved was that there are some explicit positive constants
a, b so that
logx logx
a < π(x) < b .
x x
The chapter is a modified version of an article that first appeared in Resonance, Vol. 7, No. 6,
pp. 77–87, June 2002.
37
How Far Apart are Primes? Bertrand’s Postulate
A refined version of the prime number theorem indeed implies that π(x)
has the same asymptotic expansion.
In particular, this implies that the best possible values for A and B in
Legendre’s conjecture are A = 1, B = −1.
The prime number theorem was proved independently by Hadamard and
de la Vallee Poussin. A well-known mathematician quipped once that the
proof almost immortalised these two mathematicians – they lived to be 96
and 98 respectively!
Returning to Bertrand’s postulate, after Chebychev’s first proof, other
simpler proofs appeared. A generalization of Bertrand’s postulate is
Sylvester’s theorem which was stated and used in the previous chapter.
In this chapter, we shall discuss two of the simplest proofs due to two great
minds – Ramanujan and Erdos. Most of us are told stories about Ramanu-
jan and his discoveries and it is rarely that one can find a proof of his which
is elementary enough to be actually discussed at this level. Erdos’s proof is
even more elementary and we start with it.
Erdos’s Proof
Q
We start with any natural number n and look at the product p≤n p
over all primes p ≤ n.
We shall have occasion to use the well-known and easily proved fact
asserting that the highest power to which a prime divides n! is given by the
38
How Far Apart are Primes? Bertrand’s Postulate
expression
[n/p] + [n/p2 ] + [n/p3 ] + · · · · · ·
Erdos’s proof starts with the following very beautiful observation:
Lemma. p≤n p ≤ 4n .
Q
Proof. We prove this by induction on n. It evidently holds good for small
n. Look at some n > 1 such that the result is assumed for all m ≤ n. Then,
Y Y Y
p= p p
p≤n 2p≤n+1 n+1<2p≤2n
n+1 Y
≤4 2 p
n+1<2p≤2n
then
Q weepshall use the lemma to give an upper bound for the first product
p≤n p . Here, p
e denotes the power of p dividing the middle binomial
2n
coefficient n and, the second product stands for 1 if there are no terms.
We want to see which primes p ≤ n actually contribute to 2n
.
√ n
If p2 > 2n i.e., if n ≥ p > 2n, then clearly, ep = [ 2n
p ] − 2[ n
p ] = 0 or 1.
2n
Thus, such primes divide n either to a single power or not at all.
If n ≥ 3, then a prime p ≤ n with 2n/3 < p must be at least 3 and so
p2 ≥ 3p > 2n. As 1 ≤ n/p < 3/2 for 2n/3 < p ≤ n, we have [ np ] = 1 and
[ 2n
p ] = 2 i.e.,
39
How Far Apart are Primes? Bertrand’s Postulate
2n
ep = [ 2n n
p ] − 2[ p ] = 2 − 2 = 0. Thus, these primes do not divide n when
n ≥ 3.
In other words,
√ we have:
ep ≤ 1 if 2n < p ≤ 2n/3, and ep√= 0 if 2n/3 < p ≤ n.
Finally, for the primes with p ≤ 2n, we simply take the trivial bound
pep ≤ 2n. Then, we have
Y
2n Y Y Y Y
= pep p≤ (2n) p p
n √ √
p≤n n+1<p≤2n p≤ 2n 2n<p≤2n/3 n+1<p≤2n
Y Y
≤ (2n) 42n/3 p,
√
p≤ 2n n+1<p≤2n
22n √
2n Y
≤ ≤ (2n) 2n−2 42n/3 p.
2n + 1 n
n+1<p≤2n
22n
Replacing the first term by (2n)2
, we get
Y 4n/3
p≥ √ .
(2n) 2n
n+1<p≤2n
Thus, to show that the left side has terms (i.e., that it is not 1 according to
our convention), it suffices to see whether the right hand side is bigger than
1 for all n. As usual, this will turn out to be true for large enough values of
n and will fail for small values (this only means that the inequality is good
enough for large values of n and we need to verify the original assertion
directly for the smaller values left out).
After a few trials, √
we arrive at the number n = 450 and find that
4n/3 = 4150 > (2n) 2n = (900)30 since 45 > 900.
There is nothing special about 450 excepting the fact that 2n is a per-
fect square and 450 is large enough for the inequality to hold good. This √
inequality continues to hold for n > 450 as the difference 4n/3 − (2n) 2n
is an increasing function. This last statement is simple to see by looking
at the derivative of the difference of the corresponding logarithms. Now it
is an easy exercise to verify Bertrand’s postulate for n < 450. The above
proof was essentially due to Erdos; it is a slightly simplified version of his
original argument which appears in [2].
40
How Far Apart are Primes? Bertrand’s Postulate
Ramanujan’s Proof
Let us turn to Ramanujan’s proof. It is also extremely clever and com-
pletely elementary apart from the use of what is known as Stirling’s formula
– a proof has been discussed in Resonance earlier Q[3].
In the previous proof we used an estimate for p≤n p. Here, we consider
an Padditive version of it viz., look at the so-called Chebychev function θ(x)
= p≤x log p defined for any real number x ≥ 2. One remark about why
one considers functions like the Chebychev function instead of just the
prime-counting function – weighted prime-counting is easier as the function
becomes smoother. We note two things to begin with:
Q
(i) θ(n) is simply the logarithm of p≤n p,
(ii) Bertrand’s postulate is true for a real number x if it is true for n = [x];
indeed, a prime between n and 2n is between x and 2x as well.
Let us also understand that since we are interested in primes between x
and 2x, we need a lower bound for θ(2x) − θ(x). In oher words, we need
reasonable lower as well P
as upper bounds for θ values.
Now, the expression i
i≥0 [n/p ] for the power of a prime dividing n!
gives us
X
log[x]! = Ψ(x/i),
i≥1
41
How Far Apart are Primes? Bertrand’s Postulate
42
How Far Apart are Primes? Bertrand’s Postulate
To see this, given x > 1, let us look at the area under the curve y = x1
Rx P[x] 1
between 1 and x. Evidently, this area 1 dx x = logx is less than n=1 n .
(see the figure).
So, if we consider the product p≤x (1 − p1 )−1 , then clearly,
Q
Y 1 X[x] 1
(1 − )−1 ≥ ≥ logx.
p≤x p n=1 n
where a = p (1 + p21−1 )−1 . Using ex > 1 + x for all x > 0, we get p≤x e1/p
Q Q
43
How Far Apart are Primes? Bertrand’s Postulate
defined for all n ≥ 3. Here, the sign function is the function which takes the
x
value 0 at 0 and the value |x| for any x 6= 0. Clearly, fn = 1 or 0 according
as n is prime or composite. Now, by Bertrand’s postulate, if pn ≥ 3, then
pn+1 occurs as the first prime among pn + 2, pn + 4, · · · , 2pn − 1. Therefore,
(writing pn as p for simplicity of notation),
Prime Sums
For any positive integer n, consider the set {1, 2, · · · , 2n} of the first 2n
positive integers. We claim that this set can be written as the union of n
pairs of integers {ai , bi } (1 ≤ i ≤ n) such that ai + bi is prime! Indeed,
this is clear for n = 1 as 1 + 2 = 3 is prime, and we will apply induction
on n to prove it in general. Assume that n > 1 and that our assertion
is valid for every m < n. Now, Bertrand’s postulate ensures we have a
prime p among the numbers in the set {2n + 1, 2n + 2, · · · , 4n − 1}. Writing
p = 2n + r, we have r ∈ {1, 2, · · · , 2n − 1}. Thus, note that r is odd as
p must be an odd prime. If r > 1, then by induction hypothesis, the set
{1, 2, · · · , r − 1} can be split into pairs {ai , bi } (1 ≤ i ≤ r−12 ) such that
ai + bi is prime for each i. Now, {r, r + 1, · · · , 2n} is evidently split into the
pairs {r, 2n}, {r + 1, 2n − 1}, · · · whose sums are all equal to the prime p.
Another very interesting application is the following one. By refining the
above methods, one may prove that for any positive integer k, there is
a sufficiently large N such that there is a prime between n and 2n − k
for all n > N . Applying this to k = 11, Robert Dressler showed in 1972
that every positive integer other than 1,2,4,6,9 is a sum of distinct odd
primes.
44
How Far Apart are Primes? Bertrand’s Postulate
References
[1] W H Mills, A prime-representing function, Bull. Amer. Math. Soc. 53,
pp. 604, 1947.
[2] I Niven and D Zuckermann, Introduction to number theory, John Wiley
and Sons, New York, 1960.
[3] S Ramasubramanian, Mathematical Analysis, Echoes from Resonance,
Universities Press, Hyderabad, 2001.
[4] T Apostol, Introduction to analytic number theory, Springer Interna-
tional Students Edition, Narosa Publishers, New Delhi, 1986.
[5] American Math. Monthly, 1975.
45
Sums of Powers, Bernoulli
and the Riemann Zeta function
1. Introduction
It is a beautiful discovery
Pn k due to Jakob Bernoulli that for any positive
integer k, the sum i=1 i can be evaluated in terms of, what are now
known as, Bernoulli numbers. It is said that the Bernoulli numbers were
discovered simultaneously and independently by Japanese mathematician
Seki Kowa. Seki Kowa’s discovery was posthumously published in 1712, in
his work Katsuyo Sampo. In this article, we shall discuss several methods
of evaluating the above sum. Apart from Bernoulli’s method which we
shall recall, we give a method akin to using integration, and one using
differentiation. These methods are often useful in evaluating more general
sums too as we shall indicate. We also discuss connections with the Riemenn
zeta function – some old and some new.
The chapter is a modified version of an article that first appeared in Resonance, Vol. 8, No. 7,
pp. 54–62, July 2003.
47
Sums of Powers, Bernoulli and the Riemann Zeta function
and thus
k
X k
Bk (t) = Bl tk−l .
l
l=0
In other words,
k
k k 1 X k+1k
1 + 2 + ··· + n = Bl (n + 1)k+1−l .
k+1 l
l=0
Note that it is evident from this formula that the sum of the k-th powers
of the first n natural numbers is a polynomial function of n of degree k + 1.
3. Method of ‘Integration’
For convenience, let us denote Sk (n) = 1k + 2k + · · · + nk . This is a
polynomial function of n i.e., there is a polynomial Sk (x) of degree k + 1
such that the above equality holds for all n.
The basic idea of the method we will discuss now is that (since nk = Sk (n)
− Sk (n − 1)), xk can be thought of as a ‘derivative’ of the function Sk (x). In
other words, Sk (x) itself may be thought of as an ‘integral’ of the function
xk . Of course, this is only heuristic at the moment because xk will be the
derivative of Sk at some point between x−1 and x. The correct tool to make
this precise is the ‘method of differences’ which is really a discrete analogue
of differentiation. More precisely, let us recall that the ‘backward difference’
operator is defined on any function f by (∇f )(x) = f (x) − f (x − 1) for all
x. It is trivial to see that if Pr (x) = x(x + 1) · · · (x + r − 1) for r ≥ 1 and
for all x, then (∇Pr )(x) = rPr−1 (x) for all x.
48
Sums of Powers, Bernoulli and the Riemann Zeta function
49
Sums of Powers, Bernoulli and the Riemann Zeta function
n(n + 1) n(n + 1) · · · (n + d)
f (1) + · · · + f (n) = a0 n + a1 + · · · + ad .
2 d+1
50
Sums of Powers, Bernoulli and the Riemann Zeta function
22k−1 2k
ζ(2k) = (−1)k−1 B2k π .
(2k)!
We remark in passing that the same cotangent series is the starting point
for obtaining an expression via theta series for the number of ways of writing
a positive integer as a sum of squares.
Here is a rather surprising observation. The Riemann zeta function ζ(s) is
−s
P
defined by the series n≥1 n for any complex number with Re s > 1. The
theory of the zeta function implies that its definition can be extended (not
by the same series, of course) to all values of s other than s = 1. Moreover,
the values at s and 1 − s are related by what is known as a functional
equation (thus there is the mysterious half line Re s = 1/2 in the middle
on which the Riemann hypothesis predicts all the nontrivial zeroes of ζ(s)
ought to lie). Let us now think of the naiveP idea that since ζ(k) for any
natural number k > 1 is given by the series n≥1 n−k , it is possible that
the value ζ(−k) is related to the partial sums n≤N nk . That this is indeed
P
so is a simple, beautiful observation due to J Minac [1]). Recall from the
previous discussion that there is a unique polynomial Sk (x) which coincides
with the sum 1k + · · · + nk at x = n for any natural number n and that Sk
has degree k + 1. In fact, we saw that
1 1
Bk+1 (x + 1) − Bk+1 (1)
Z Z
Bk+1
Sk (x − 1)dx = = (−1)k .
0 0 k+1 k+1
We claim: Z 1
Bk+1
ζ(−k) = Sk (x − 1)dx = (−1)k .
0 k+1
Actually, one can use the functional equation for the zeta function to
conclude this but we follow a more elementary method of obtaining analytic
continuation of the zeta function which will also prove this claim.
51
Sums of Powers, Bernoulli and the Riemann Zeta function
52
Sums of Powers, Bernoulli and the Riemann Zeta function
The infinite sum on the right hand side converges for Re (s) > −m and
thus we have an expression for ζ(s) for such s. At this point, we evaluate
it at s = 1 − m. Rather surprisingly, this pretty but simple idea does not
seem to have been thought of until very recently when it was done so by
Ram Murty and M Reece. We get
m−1
(−1)m
1 q m−1 1
X
ζ(1 − m) = 1 − + − (−1) (ζ(1 − m + q) − 1).
m m(m + 1) q q+1
q=1
Therefore, we get
k Z 1
(−1)k+1
X k+1
(−1)r Sk−r (x − 1)dx = .
r+1 0 k+2
r=0
References
[1] J Minac, Expo. Math., Vol. 12, pp. 459–462, 1994.
53
Frobenius and His Density Theorem
for Primes
1. Introduction
Our starting point is the following problem which appeared in the recent
IMO (International Mathematical Olympiad):
If p is a prime number, show that there is another prime number q such
that np − p is not a multiple of q for any natural number n.
Now, this problem itself can be solved using elementary mathematics
(otherwise, it would not be posed in the IMO). However, how does one
guess that such a thing ought to be true? Can we produce an abundance of
such problems in some systematic manner? We take this problem as a point
of reference to discuss some deep number theory (which is already a century
old) which not only solves this problem, but also gives us an understanding
of why such facts are true and what more one can expect. The main theorem
under discussion is known as Frobenius’s density theorem.
55
Frobenius and His Density Theorem for Primes
56
Frobenius and His Density Theorem for Primes
57
Frobenius and His Density Theorem for Primes
The second and third equalities above show that P (X) is reducible mod-
ulo any q n and any pn respectively. Also since p ≡ 1 mod 8, p is a quadratic
residue modulo 2 and, therefore, modulo any 2n ; the second equality above
again shows that P (X) is the difference of two squares modulo 2n , and
hence reducible mod 2n .
If ` is a prime 6= 2, p, q, at least one of ( p` ), ( q` ) and ( pq
` ) is 1 by the product
formula ( ` ) · ( ` ) · ( ` ) = 1 that we noted earlier. According as ( p` ), ( q` ) or
p q pq
( pq
` ) = 1, the second, third or fourth equality shows that P (X) is reducible
mod `n for any n.
We mention, in passing, a very simple but important general method of
proving the irreducibility of an integral polynomial. This will also set up
the notation for our main statement when we address (∗). To illustrate it,
consider the polynomial p(X) = X 4 + 3X 2 + 7X + 4. Modulo 2, we have
p(X) = X(X 3 + X + 1) and both factors are irreducible over the field
Z/2Z. We say that the decomposition type of p(X) mod 2 is 1, 3. Therefore,
either p is irreducible over Z or if not, it is a product of a linear factor
and an irreducible factor of degree 3 over Z. But, modulo 11, we have
p(X) = (X 2 + 5X − 1)(X 2 − 5X − 4) where both factors are irreducible
over the field Z/11Z. That is, the decomposition type of p mod 11 is 2, 2.
Thus, it cannot be that p has a linear factor over Z. In other words, p must
be irreducible over Z.
58
Frobenius and His Density Theorem for Primes
59
Frobenius and His Density Theorem for Primes
what the word ‘effective’ means here). Chebotarev’s idea of proving this
has been described by two prominent mathematicians as “a spark from
heaven”. In fact, this theorem was proved in 1922 (“while carrying water
from the lower part of town to the higher part, or buckets of cabbages to
the market, which my mother sold to feed the entire family”) and Emil
Artin wrote to Hasse in 1925: “Did you read Chebotarev’s paper? ... If
it is correct, then one surely has the general abelian reciprocity laws in
one’s pocket...” Artin found the proof of the general reciprocity law in 1927
using Chebotarev’s technique (he had already boldly published the reci-
procity law in 1923 but admitted that he had no proof). Nowadays, Artin’s
reciprocity law is proved in some other way and Chebotarev’s theorem is
deduced from it!
To state Chebotarev’s theorem, we recall one notion – the Frobenius
map. The idea is that given a monic integral polynomial f and its split-
ting field K, we can associate to any prime p 6 |disc(f ), an element Φp
of Gal(f ) in a natural manner. If we can do this, one may expect that
the decomposition type of f modulo p coincides with the cycle pattern
of Φp . It can almost be done except that a prime p gives rise to a con-
jugacy class of elements in Gal(f ). We do not define the Frobenius con-
jugacy class here as it is somewhat technical and merely explain some
properties it has. For any prime number p, the p-th power map F robp
is an automorphism of the field F¯p which is identity on Fp . Therefore,
F robp permutes the roots of any polynomial over Fp . Indeed, the Ga-
lois theory of finite fields amounts to the statement that if g is a poly-
nomial over Fp with simple roots, then the cycle pattern of F robp viewed
as a permutation of the roots of g coincides with the decomposition type
of g over Fp . In our case, we start with an integral polynomial f and
look at it modulo p for various primes p. The basic theory of algebraic
numbers shows that whenever p 6 |disc(f ), the automorphism F robp gives
rise to a conjugacy class in Gal(f ), called the Frobenius conjugacy class
modulo p.
In Frobenius’s density theorem, one cannot distinguish between two primes
p, q defining different conjugacy classes C(x) and C(y) but some powers of
x and y are conjugate. For instance, for the polynomial X 10 −1, the decom-
position type modulo primes congruent to 1, 3, 7, 9 mod 10 are, respectively,
1, 1, 1, 1, 1, 1, 1, 1, 1, 1; 1, 1, 4, 4; 1, 1, 4, 4; 1, 1, 2, 2, 2, 2.
Frobenius’s theorem cannot distinguish between primes which are 3 mod
10 and those which are 7 mod 10; they define different conjugacy classes
in Gal(X 10 − 1). Thus, it would imply that the number of primes ≡ 3 or 7
mod 10 is infinite but doesn’t say whether each congruence class contains
infinitely many primes. This is what Chebotarev’s theorem asserts.
60
Frobenius and His Density Theorem for Primes
References
[1] D Berend and Y Bilu, Polynomials with roots modulo every integer,
Proc. Amer. Math. Soc., Vol. 124, pp. 1663–1671, 1996.
[2] P Stevenhagen and H W Lenstra, Jr., Chebotarev and his density
theorem, Mathematical Intelligencer, Vol. 18, pp. 26–37, 1996.
61
When is a Decimal Expansion
Irrational?
√
Everyone learns in school that 2 is irrational. This, along with Euclid’s
proof of infinitude of primes, is probably the first time she encounters√a
proof by contradiction. Most students know in school that the value of 2
is approximately 1.414 but, more often than not, this aspect is not pursued
further in detail. The decimal 0.999 · · · where 9 recurs indefinitely is under-
stood (after some persuasion perhaps) to be none else than the number 1.
The problem here is that the concept of limit takes some time to sink in.
Given that start, they can see easily that numbers like 0.142857 where
the digits overlined recur indefinitely, are rational. Even if the recurring
string occurs after an initial string (for example, a decimal expansion like
27.142857), it still gives us a rational value only because it is the sum of a
geometric series with a ratio of the form 10k .
It is not hard to prove that this is a necessary condition as well, that
is, a decimal expansion of a real number represents a rational number if,
and only if, after the decimal place, there is a finite (possibly empty) string
after which the digits consist of a finite string (possibly consisting entirely
of zeroes) recurring indefinitely.
Thus, for instance, the number 0.101001000 · · · where the number of
zeroes keeps increasing by 1 has to represent an irrational number. However,
from the decimal expansion, sometimes it is not clear whether there is such
an eventual recurrence or not. This could be due to our present state of
knowledge. For instance, one could define a number 0.1010 · · · where the
number of zeroes occurring at the n-th step is either increased by one or
n
kept the same according as to whether the number 22 + 1 is prime or not.
Since one does not know whether there are infinitely many such primes, we
cannot say at present as to whether the above decimal represents a rational
or an irrational number.
The decimal 0.1234 · · · , where the natural numbers are written in se-
quence, is clearly irrational since, for instance, the number of zeroes occur-
ring in the powers of 10 keeps increasing. The decimal 0.235711 · · · , where
the set of primes is written down in sequence, is also irrational. This is
because there is a prime of the form 10n a + 1 for an arbitrary n – this
was proved in an earlier chapter. Here is another elementary general result
(see [1]).
The chapter is a modified version of an article that first appeared in Resonance, Vol. 9, No. 3,
pp. 78–80, March 2004.
63
When is a Decimal Expansion Irrational?
References
[1] A McD Mercer, American Mathematical Monthly, Vol. 101,
pp. 567–568, 1994.
64
Revisiting Kummer’s
and Legendre’s Formulae
Legendre’s Formula
Let p be a prime number and let ak · · · a1 a0 be the base-p expansion of
a natural number n. We shall show that if Legendre’s formula
n − ki=0 ai
P
n − s(n)
vp (n!) = = (2)
p−1 p−1
The chapter is a modified version of an article that first appeared in Resonance, Vol. 10, No. 2,
pp. 62–71, February 2005.
65
Revisiting Kummer’s and Legendre’s Formulae
holds good for n, then it also holds good for pn + r for any 0 ≤ r < p. Note
that the base-p expansion of pn + r is
ak · · · a1 a0 r.
m−s(m)
Let us denote, for convenience, the number p−1 by f (m) for any
natural number m. Evidently,
pn − ki=0 ai
P
f (pn + r) = = n + f (n).
p−1
On the other hand, it follows by induction on n that
Since it is evident that f (m) = 0 = vp (m!) for all m < p, it follows that
f (n) = vp (n!) for all n. This proves Legendre’s formula.
Note also that the formula
n n n
vp (n!) = + 2 + 3 + ···
p p p
Kummer’s Algorithm
As before p is any prime number. For any natural numbers r and s, let
us denote by g(r, s) the number of ‘carry-overs’ when the base-p expansions
of r and s are added. Kummer’s result is that for k ≤ n,
n
vp = g(k, n − k). (4)
k
Once again, this is clear if n < p, as both sides are then zero. We shall
show that if the formula holds good for n (and every k ≤ n), it does so for
pn + r for 0 ≤ r < p (and any k ≤ pn + r). This would prove the result for
all natural numbers.
pn+r
Consider any binomial coefficient pm+a for 0 ≤ a < p.
First, suppose a ≤ r.
66
Revisiting Kummer’s and Legendre’s Formulae
pm + a = bk · · · b0 a
p(n − m) + (r − a) = ck · · · c0 r − a.
a ≤ r.
Thus, we are through in the case when
pn + r
Now suppose that r < a. Then vp is equal to
pm + a
n − 1 = ak · · · ad+1 ad − 1 p − 1 · · · p − 1.
67
Revisiting Kummer’s and Legendre’s Formulae
pm + a = bk · · · b0 a,
p(n − m − 1) + (p + r − a) = ck · · · c0 p + r − a.
We add these using that fact that there is a carry-over in the beginning
and that 1 + bi + ci = p for i < d. Since there is a carry-over at the first
step as well as at the next d steps, we have
pn + r = ∗ ∗ · · · ad 0 · · · 0 r
References
[1] R Honsberger, In Polya’s Footsteps, published and distributed by the
Mathematical Association of America, 1997.
[2] P Ribenboim, The Book of Prime Number Records, Springer-Verlag,
1996.
68
Bessels Contain Continued Fractions
of Progressions
1. Introduction
The January 2000 issue of Resonance carried a nice article on continued
fractions by Shailesh Shirali. After discussing various continued fractions
for numbers related to e, he left us with the intriguing question as to how
one could possibly evaluate the continued fraction
1 1 1
········· .
1+ 2+ 3+
The question is interesting because this continued fraction is simpler-
looking than the ones which were studied in that article. We answer this
question here and show that the discussion naturally involves the Bessel
functions, thus explaining the title. However, we shall begin with some
details about continued fractions which complement his discussion. One
place where continued fractions are known to appear naturally is in the
study of the so-erroneously-called Pell’s equation.
In a series of very interesting articles in Resonance, Amartya Kumar
Dutta had discussed various aspects of Mathematics in ancient India. In
particular, he discussed Brahmagupta’s and Bhaskara’s work on Samasab-
havana and the Chakravala method for finding solutions to ‘Pell’s equation’.
In fact, it is amusing to recall what Andre Weil, one of the great mathe-
maticians of the last century wrote once, while discussing Fermat’s writings
on the problem of finding integer solutions to x2 − Dy 2 = 1:
The chapter is a modified version of an article that first appeared in Resonance, Vol. 10, No. 3,
pp. 80–87, March 2005.
69
Bessels Contain Continued Fractions of Progressions
for some integer t. This gives all solutions, and the smallest solution in
natural numbers x, y is obtained by taking t = −1 and turns out to be
(185, 96).
The reader is left with deriving similarly the corresponding expression
for any linear equation.
70
Bessels Contain Continued Fractions of Progressions
√
(1 + 5)/2. This is because the value s satisfies p s = 1 + 1/s and is posi-
tive. Similarly, the SCF [1 ; 3, 2, 3, 2, · · · ] = 5/3, as it gives the√quadratic
equation s − 1 = (s + 1)/(3s + 4), and [0 ; 3, 2, 1, 3, 2, 1, · · · ] = ( 37 − 4)/7
as it gives the equation s = (3 + 2s)/(10 + 7s), etc.
Consider a quadratic Diophantine equation in two variables
For any integral solution, the expression inside the square root (which we
write as ry 2 + 2sy + t now) must be a perfect square, say v 2 . Once again,
solving this as a polynomial in y, we get
p
ry + s = ± (s2 − rt + rv 2 ).
x2 − ry 2 = −1, x2 − ry 2 = 1.
71
Bessels Contain Continued Fractions of Progressions
For example, √
13 = [3 ; 1, 1, 1, 1, 6, · · · ].
The period is 5 which is odd. The penultimate convergent to the first
period is
1 1 1 1 18
3+ = .
1+ 1+ 1+ 1 5
Therefore, (18, 5) is a solution of u2 − 13v 2 = −1.
The penultimate convergent to the second period is computed to be
649/180. Therefore, (649, 180) is a solution of u2 − 13v 2 = 1.
e+1 e2 + 1
[2 ; 6, 10, 14, · · · ] = , [1 ; 3, 5, 7, · · · ] = .
e−1 e2 − 1
The SCF’s here involve terms in arithmetic progression. What about
a general SCF of the form [a ; a + d, a + 2d, · · · ]? For example, can the
SCF [0 ; 1, 2, 3, · · · ] be evaluated in terms of some ‘known’ numbers and
functions? Shirali started with the differential equation (1 − x)y 00 = 2y 0 + y
which he remarked “does not seem to solvable in closed form”.
We start with any arithmetic progression a, a + d, a + 2d, · · · where a is
any real number and d is any non-zero real number, and show how it can
be evaluated.
Let us consider the differential equation
dxy 00 + ay 0 = y. (5)
Actually, heuristic reasons can be given as to why one looks at this differ-
ential equation but we directly start with it here and show its relation to our
problem. Let y = y(x) be a solution of the above differential equation satis-
fying y(0) = ay 0 (0). Let us denote the r-th derivative of y by yr for simplicity
of notation. By repeated differentiation, we get dxyr+2 + (a + rd)yr+1 = yr
for all r ≥ 0 (with y0 denoting y). Therefore, we have
y0 dxy2 dx dx
=a+ =a+ ······ . (6)
y1 y1 a + d+ a + 2d+
Observe that
y(1/d)
[a ; a + d, a + 2d, a + 3d, · · · ] = .
y 0 (1/d)
72
Bessels Contain Continued Fractions of Progressions
these are usually referred to as Bessel functions of the first kind. Here Γ(s)
is the Gamma function. If α is not an integer, then J−α (defined in the
obvious manner) is another independent solution to the Bessel differential
equation above. Closely related to the Jα is the so-called modified Bessel
function of the first kind
X (x/2)2n+α
Iα (x) = . (11)
n!Γ(n + 1 + α)
n≥0
Thus, we have
1 1 1 I1 (2)
········· = .
1+ 2+ 3+ I0 (2)
The function Iα (x) is a solution of the differential equation x2 y 00 + xy 0 −
(x2 + α2 )y = 0. Indeed, Iα (x) = i−α Jα (ix) for each x. Using the relation
√
Γ(s + 1) = sΓ(s) and the value Γ(1/2) = π, it is easy to see that the
solution function
73
Bessels Contain Continued Fractions of Progressions
X xn+1
y = c0 + c0 (12)
(n + 1)!a(a + d) · · · (a + nd)
n≥1
above, is related to the modified Bessel function of the first kind as:
In particular,
Conclusion
Before finishing, we recall some SCF’s evaluated out by Shirali:
e+1 e2 + 1
[2 ; 6, 10, 14, · · · ] = , [1 ; 3, 5, 7, · · · ] = .
e−1 e2 − 1
Our formula above yields for the same SCF’s the expressions:
I−1/2 (1/2)
[2 ; 6, 10, 14, · · · ] = , (15)
I1/2 (1/2)
I−1/2 (1)
[1 ; 3, 5, 7, · · · ] = . (16)
I1/2 (1)
Therefore, for these special parameters, the value of the modified Bessel
function is expressible in terms of e and one can recover Shirali’s expres-
sions.
74
The Prime Ordeal
Prime numbers have fascinated mankind through the ages. In fact, one
may think that we know all about them. However, this is not so! One
does not know the answers to many basic questions on primes. We shall
concentrate here mainly on questions and discoveries whose statements are
elementary and accessible. Right at the end, we mention a result whose
statement is simple but whose proof uses rather sophisticated mathematics.
Even here, we do not try to be exhaustive. The subject is too vast for that
to be possible.
1. Introduction
Let us start with the first major discovery about primes, which is the
proof by Euclid’s school that there are infinitely many prime numbers.
Euclid’s proof of the infinitude of primes will eternally remain beautiful no
matter what advances modern mathematics makes. In spite of its simplicity,
it still retains quite a bit of mystery. For instance, it is unknown as yet
whether the product of the first few primes added to 1 takes a prime value
infinitely often. It is even unknown whether it takes a composite value
infinitely often! Do you see the mystery? What is the first time we get
some composite number? Does anyone know the answer already? Anyway,
The chapter is a modified version of an article that first appeared in Resonance, Vol. 13, No. 9,
pp. 866–881, September 2008.
75
The Prime Ordeal
76
The Prime Ordeal
77
The Prime Ordeal
78
The Prime Ordeal
Giuseppe Giuga conjectured in 1950 that this characterises primes; that is,
Conjecture (Giuga 1950): n−1 n−1 ≡ −1 mod n ⇒ n is prime.
P
k=1 k
As he showed, the conjecture can be reformulated as follows:
Theorem. n−1 n−1 ≡ −1 mod n if, and only if, for each prime divisor
P
k=1 k
p of n, both p and p − 1 divide np − 1.
Pn−1 n−1
Equivalently, a composite number n satisfies k=1 Pk ≡Q −1 mod n
if, and only if, it is a Carmichael number such that p|n p1 − p|n p1 is a
natural number.
In the above statement, the sum and the product run over primes and
p|n denotes ‘p divides n’.
Pp−1 r
Proof. Note that for any prime p, we have k=1 k ≡ −1 or 0 mod p
according as whether p − 1 divides r or not.
Therefore, for a prime p dividing n, we have
n−1
X p−1
X 2p−1
X n−1
X
n−1 n−1 n−1
k ≡ k + k + ··· + k n−1
k=1 k=1 k=p+1 k=n−p+1
79
The Prime Ordeal
Remarks. A composite number n such that p|( np − 1) for each prime p|n,
is called a Giuga number. Equivalently, p|n p1 − p|n p1 ∈ N. Then, Giuga’s
P Q
conjecture amounts to the assertion that there is no Giuga number which is
also a Carmichael number. As of today, only 12 Giuga numbers are known
and all of them have sum minus product (of reciprocals of prime divisors)
equal to 1. The numbers 30, 858, 1722 are Giuga numbers. Until now, no
odd Giuga numbers have been found. Any possible odd Giuga number must
have at least 10 prime factors because the sum 13 + 15 + 71 + 11
1 1
+ 13 1
+ 17 +
1 1 1
23 + 29 + 31 < 1.
In an article in Volume 103 of the American Mathematical Monthly of
1996, David Borwein, Jonathan Borwein, Peter Borwein and Roland Gir-
gensohn propose that a good way to approach Giuga’s conjecture is to study
Giuga numbers in general. More generally, they define a Giuga sequence to
be a finite Qsequence n1 < n2 < · · · < nr of natural numbers such that
P r 1 r 1
i=1 ni − i=1 ni is a natural number. Thus, a Giuga sequence consist-
ing of primes gives rise to a Giuga number, viz., to the product of those
primes. The smallest Giuga sequence where the sum minus product is > 1,
has 59 factors! Here is an easy method to produce arbitrarily long Giuga
sequences.
Theorem. Suppose n1 < n2 < · · · < nr is a Giuga sequence satisfying
nr = r−1
Q
i=1 n i − 1. Then, the sequence n1 < n2 < · · · < ñr , ñr+1 is a Giuga
Qr−1
sequence whoseQr−1 sum minus product is the same, where ñ r = i=1 ni +
1, ñr+1 = ñr i=1 ni − 1.
Starting with a sequence like 2, 3, 5 say, this gives Giuga sequences of
arbitrary lengths whose sum minus product is 1. The proof is a simple
exercise of manipulation. In fact, one has the following nice result:
80
The Prime Ordeal
81
The Prime Ordeal
4. All’s Bell
In this section, we discuss a conjecture due to Djuro Kurepa which can
be stated in elementary language but the proof which appeared last year
involves some sophisticated mathematics. Those who have learnt Galois
theory would be able to appreciate it but others can also get a flow of the
argument. Of course, the fact that an elementary statement may require
very sophisticated methods should not come as a surprise. A case in point
is Fermat’s last theorem (FLT) which says that for an odd prime p, there
do not exist nonzero integers x, y, z such that xp +y p +z p = 0. The question
of Kurepa doesn’t quite require the kind of sophisticated mathematics re-
quired in FLT though.
Pp−1Kurepa conjectured in 1971 that for any odd prime
p, the sum Kp := n=0 n! is not a multiple of p. Of course K2 = 2. This
is, of course, not a characterisation of primes; for example, K4 = 10. The
proof (only in 2004) of Kurepa’s conjecture due to D Barsky and B Ben-
zaghou involves the so-called Bell numbers named after Eric Temple Bell.
One way of defining the Bell numbers is as follows. The n-th Bell number
Pn is the number of ways of writing an n-element set as s union of non-
empty subsets. We see that P1 = 1, P2 = 2, P3 = 5, P4 = 15, P5 = 52 etc.
There is a lot of combinatorics involving the Bell numbers.
Pn From combi-
n
natorial considerations, one can prove that Pn+1 = k=0 k Pk , where we
have written P0 to stand for 1. From this, it is easy to prove (analogously
to the proof for Bernoulli numbers) that the generating function for Pn ’s is
given by
∞ ∞
X X xn
F (x) = Pn xn = · · · · · · (♠).
(1 − x)(1 − 2x) · · · (1 − nx)
n=0 n=0
The Kurepa question can be formulated in terms of the Bell numbers P eas-
ily. It turns out using some elementary combinatorics that Pp−1 ≡ p−2 n=0 n!
modulo p. Thus, since Kp is the sum of (p − 1)! with the right hand side
above, Kurepa’s conjecture amounts to the statement that Pp−1 6≡ 1 mod-
ulo p because (p − 1)! ≡ −1 modulo p. The idea of the proof Kurepa’s
conjecture is to consider what is known as the Artin–Schreier extension
Fp [θ] of the field Fp of p elements, where θ is a root (in the algebraic clo-
sure of Fp ) of the polynomial xp − x − 1. This is a cyclic Galois extension
of degree p over Fp . Note that the other roots of xp − x − 1 are θ + i for
i = 1, 2, · · · , p − 1. The theory of such extensions is named after Emil Artin
and Otto Schreier. The reason this field extension comes up naturally is
as follows. The generating series F (x) of the Bell numbers can be evalu-
ated modulo p; this means one computes a ‘simpler’ series Fp (x) such that
F (x) − Fp (x) has all coefficients multiples of p. Since Kurepa’s conjecture
is about the Bell numbers Pp−1 considered modulo p, it makes sense to
82
The Prime Ordeal
consider Fp (x) rather than F (x). Reading the equality (♠) modulo p, one
gets
p−1 X
X xip+n
Fp (x) =
(1 − x) · · · (1 − (ip + n)x)
n=0 i≥0
p−1 X
X xn xip
=
1 − (ip + 1)x) · · · (1 − (ip + n)x) (1 − x) · · · (1 − ipx)
n=0 i≥0
p−1 X i
xn xp
X
≡
1 − (ip + 1)x) · · · (1 − (ip + n)x) (1 − x) · · · (1 − px)
n=0 i≥0
modulo p. Therefore,
Pp−1 n (1
n=0 x − (n + 1)x) · · · (1 − (p − 1)x)
Fp (x) =
1 − xp−1 − xp
Pn ≡ −T r(θcp )T r(θn−cp −1 ),
83
The Prime Ordeal
6. Sundries
We will finish with a few more remarks about primes. We mentioned
Bouniakowsky’s conjecture which asserts the infinitude of prime values.
Can a polynomial take only prime values? It is again an easy, elementary
exercise to prove that there is no nonconstant polynomial in some vari-
ables x1 , · · · , xr which takes only prime values at all integers. However, it
is a deep consequence of the solution of Hilbert’s 10th problem by Hilary
Putnam, Martin Davis, Julia Robinson and Yuri Matiyashevich that there
exist polynomials f (x1 , · · · , xr ) over integers such that the set of positive
values taken by f equals the set of prime numbers! Of course, the polyno-
mials does take negative values as well as take certain prime values more
than once. Indeed, one can take f to be of degree 25 and r to be 26. This
expresses the fact that the set of prime numbers is a Diophantine set.
84
The Prime Ordeal
Using this, and nothing more than the Chinese remainder theorem, they
showed that any prime n can be proved to be prime by expressing it as
n = N1 + N2 + · · · + Nk where p1 , · · · , pk are the first k primes and n is not
divisible by any of them while each Ni is divisible by all the pj with j 6= i
and not by pi .
85
Extending Given Digits to Make Primes
or Perfect Powers
1. Introduction
Start with any string of digits. Can we always put down some more digits
on the right of it to get a prime? Can we similarly get a power of 2? How
about a power of 3? It turns out that the answers to all these questions are
in the affirmative. Our discussion will be elementary excepting a concrete
consequence of a weak version of the prime number theorem. For a really
detailed analysis of the proportion of primes with given starting digits, the
interested reader is referred to look at ergodic theory. There is no prime-
producing polynomial in a single variable – this is trivial to see. However,
it turns out (as other concrete consequences of the properties of primes
like Bertrand’s postulate and the prime number theorem) that there are
exponential type of functions which produce infinitely many primes. One
n
such is the sequence of integer parts of t3 for a certain positive real number
s
..
t. Another is the sequence of integer parts of 22 for a certain real s > 0.
We shall prove these also.
87
Extending Given Digits to Make Primes or Perfect Powers
Proof. For any n, consider the intervals [0, 1/n), [1/n, 2/n), · · · , [(n −
1)/n, 1). By the pigeon-hole principle, among the fractional parts of
θ, 2θ, · · · , nθ, (n + 1)θ, there must be at least two (say rθ, sθ with r <
s ≤ n + 1) which are in the same interval [(k − 1)/n, k/n). But then the
fractional part of (s − r)θ lies in [0, 1/n). Therefore, each [(m − 1)/n, m/n]
contains the fractional part of some dθ where d is a multiple of s−r. As every
subinterval (x, y) of (0, 1) contains an interval of the form [(m − 1)/n, m/n]
for some m, n, the claim asserted follows.
Using this, one may extend any given digits to produce powers as follows.
Lemma 1. Let a > 1 be not a power of 10 and let A be any given natural
number. Then, one may add digits to the right end of the digits of A to
obtain some power of a.
Proof. For any a as above, log10 (a) is irrational because, if it is u/v, then
10u = av which implies by uniqueness of prime decomposition that a must
be a power of 10, a contradiction of the hypothesis. So, it follows by the
above observation that each interval (x, y) ⊆ (0, 1) contains some fractional
part {n log10 (a)}. Now, suppose A has d+1 digits; that is, 10d ≤ A < 10d+1 .
Then, we consider x ∈ (0, 1) such that 10x = A/10d . Choosing some large
n so that 101/n < 1 + A1 (as limn→∞ 101/n = 1), we consider the point y ∈
1/n
(x, 1) with 10y = 1010d A . Note that y < 1 because 101/n A < A + 1 ≤ 10d+1 .
If the fractional part {r log10 (a)} ∈ (x, y), we have
x < r log10 (a) − k < y
for some positive integer k. Taking 10-th powers, we have
ar
10x < < 10y
10k
which gives
10k−d A < ar < 10k−d 101/n A < 10k−d A + 10k−d .
Thus, ar has been obtained by adding k − d digits to the right of the base
10 expansion of A.
Illustration. Let us see how to demonstrate the above lemma for a small
number. Let us begin with A = 4 and a = 2. Of course A itself is a
power of 2 but let us see what we get from the above lemma. In the above
notation, d = 0 and x = log10 (4). The choice n = 16 is large enough so that
101/n < 1 + 1/4 = 1.25. Then y = x + 1/16 and the choice of r, k such that
2r 1
10x = 4 < k
< 10y = 4 +
10 16
can be taken to be r = 12, k = 3. Hence 4 can be extended to 212 = 4096.
88
Extending Given Digits to Make Primes or Perfect Powers
89
Extending Given Digits to Make Primes or Perfect Powers
This means bn < bn+1 < cn+1 < cn which ensures that the sequence {bn }
converges to some real number s as n → ∞. Notice that for this number s,
s
2 ..
the sequence an = 22 satisfies pn < an < pn + 1. Hence pn = [an ]. This
completes the proof.
Remarks.
(i) The above formula is not a practical one. Since there is a choice of pn ’s
allowed, the real number s is not unique. One possible value of s is
1.9287800 · · · and the primes pn defined by the lemma grow much too
fast. For example, p4 has 5000 digits.
(ii) A result earlier to Wright’s result above (in fact, the result which mo-
tivated Wright’s theorem) is due to W H Mills [2] in 1947. This uses a
result on primes which is considerably deeper than Bertrand’s postu-
late. This deeper result alluded to is due to the British mathematician
A E Ingham; he derived in 1937 (see [3]) the following concrete conse-
quence of the prime number theorem:
5/8
There is a positive number c such that pn + cpn > pn+1 for all n, where
p1 < p2 < p3 < · · · is the sequence of all primes.
n
Lemma 4. There exists a real number t > 0 such that [t3 ] is a prime for
every n.
Proof. Start with Ingham’s result and choose a large N > c8 . Look at
the prime pn such that pn < N 3 < pn+1 . Then, we have
90
Extending Given Digits to Make Primes or Perfect Powers
Indeed, the inequality vn > vn+1 is simply the second inequality in ♥; the
inequality un+1 > un is the first inequality of ♥ and the inequality vn+1 >
un+1 is obvious. Hence, the sequence {un } is a bounded, monotonically
increasing sequence and must have a limit t. Clearly, t = limn→∞ un satisfies
un ≤ t < vn . Thus,
n
prn ≤ t3 < prn + 1.
n
This proves that [t3 ] = prn for all n.
References
[1] E M Wright, A prime-representing function, Amer. Math. Monthly,
Vol. 58, No. 9, pp. 616–618, 1951.
[2] W H Mills, A prime-representing function, Bull. Amer. Math. Soc., 53,
pp. 604, 1947.
[3] A E Ingham, On the difference between consecutive primes, Quart.
J. Math. Oxford Ser., Vol. 8, pp. 255–266, 1937.
91
An Irrational Walk and Why 1
is Not Congruent
1. Introduction
The subject of Diophantine equations is an area of mathematics where
solutions to very similar-looking problems can vary from the elementary to
the deep. Problems are often easy to state, but it is usually far from clear
whether a given one is trivial to solve or whether it must involve deep ideas.
Fermat showed that the equations X 4 + Y 4 = Z 4 and X 4 − Y 4 = Z 4 do
not have nontrivial solutions in integers. He did this through a method now
known as the method of descent. In fact, he discovered this while working
on a Diophantine problem called the congruent number which we discuss
below. There are other situations where these equations arise naturally.
One such problem we discuss is the following.
#
#
#
#
#
The chapter is a modified version of an article that first appeared in Resonance, Vol. 17, No. 1,
pp. 76–82, January 2012.
93
An Irrational Walk and Why 1 is Not Congruent
94
An Irrational Walk and Why 1 is Not Congruent
obtained and progress has been made which is likely to lead to its complete
solution.
The rephrasing in terms of arithmetic progressions of squares emphasizes
a connection of the problem with rational solutions of the equation y 2 =
x3 − d2 x.
Such equations define elliptic curves.
It turns out that:
d is a congruent number if, and only if, the elliptic curve Ed : y 2 =
x3 − d2 x has a solution with y 6= 0.
In fact, a2 + b2 = c2 , 12 ab = d implies bd/(c − a), 2d2 /(c − a) is a rational
solution of y 2 = x3 − d2 x.
Conversely, a rational solution of y 2 = x3 − d2 x with y 6= 0 gives the
rational, right-angled triangle with sides (x2 − d2 )/y, 2xd/y, (x2 + d2 )/y
and area d.
In a nutshell, here is the reason we got this elliptic curve. The real solu-
tions of the equation a2 + b2 = c2 defines a surface in 3-space and so do the
real solutions of 21 ab = d. The intersection of these two surfaces is a curve
whose equation in suitable co-ordinates is the above curve.
#
#
#
#
#
Recall that the rule is to walk on a straight line to some point of the
middle vertical line as in the figure and, on reaching that point, walk to-
wards the opposite corner along a straight line. Thus, we have a path as
in the figure consisting of one segment of length r until the middle line is
reached and the other of length s from that point to the opposite corner.
The question is whether we can follow such a path with both the distances
r, s rational numbers. Suppose such a ‘rational’ path is possible. Let us call
the vertical distance x on the middle line from the bottom to the point we
95
An Irrational Walk and Why 1 is Not Congruent
96
An Irrational Walk and Why 1 is Not Congruent
In other words,
(p2 − u2 )2
4Q4 − (p2 + u2 )Q2 + = 0.
8
This means that the discriminant (p2 +u2 )2 −2(p2 −u2 )2 = 6p2 u2 −p4 −u4
must be a perfect square, say d2 . But then the general algebraic identity
97
Covering the Integers
{n : fb(n) 6= 0},
{n : f (n) 6= 0},
of f is infinite.
Suppose f is non-zero and that {n : f (n) 6= 0} is finite. Then,
Proof. P
F (z) := n≥1 fb(n)z n is a non-zero polynomial. But, if M = M ax(|f (n)| :
n ≥ 1), then for |z| < 1, we have
XX XX
|F (z)| = | ( f (d))z n | ≤ |f (d)||z|n
n d|n n d|n
XX X m|z|
≤ ( M )|z|n ≤ M n|z|n = < ∞.
n n
1 − |z|2
d|n
99
Covering the Integers
P
Note that µ b(n) = d|n µ(d) = 0 for n > 1. So, µ b has finite support.
So, if there were only finitely many primes, µ would have finite support
contradicting the ‘uncertainty’ result proved above.
The above type of argument has interesting applications to what are
known as covering congruences. These can be described as follows.
Just as every natural number is either odd or even, one can see that for
any k, the congruences
x ≡ 1 (mod 2), x ≡ 1 (mod 4), · · · , x ≡ 1 (mod 2k ), x ≡ 0 (mod 2k )
cover the set of all integers. In other words, every integer satisfies at least
one of these congruences. Similarly, for any positive integer N , the set of
congruences x ≡ i (mod N ) for i = 0, 1, · · · , N − 1 covers the integers.
Thus, these are called sets of ‘covering congruences’. The general defini-
tion is the following.
Let a1 , · · · , ak be integers and let n1 , · · · , nk be positive integers. The
set of congruences x ≡ ai (mod Sk ni ) for i = 1, · · · , k is called a set of
covering congruences if Z = i=1 (ai + ni Z). We write the set in short as
a1 (n1 ), a2 (n2 ), · · · ,
ak (nk ).
Note that in both the cases P above, the ‘moduli’ n1 , · · · , nk of the congru-
ences have the property that ki=1 n1i = 1.
A set of covering congruences a1 (n1 ), · · · , ak (nk ) is called a disjoint cov-
ering system if every integer satisfies exactly one of the congruences x ≡ ai
(mod ni ).
Note that the two sets of covering congruences above, viz.,
and
0(N ), 1(N ), · · · , N − 1(N )
are disjoint covering systems. Note also that at least two of the moduli are
the same. That this must always be true is the assertion of the following
proposition whose proof follows the method we started the article with.
PROPOSITION
Let a1 (n1 ), a2 (n2 ), · · · , ak (nk ) be a disjoint system of covering
P congru-
ences where n1 ≤ n2 ≤ · · · ≤ nk . Then, nk−1 = nk . Moreover kj=1 n1j = 1.
Proof. We may assume without loss of generality that 0 ≤ aj < nj for
each j ≤ k. By the hypothesis, we have for |z| < 1,
k k ∞
X z aj XX
aj +rnj
X 1
n = z = zn = .
1−z j 1−z
j=1 j=1 r=0 n≥0
100
Covering the Integers
Thus, this function has a pole at the point z = 0. But, if ni < nk for all
i < k, then this function has a pole at z = e2iπ/nk , a contradiction. This
proves the first assertion.
Let n = LCM (n1 , · · · , nk ). Then, we have
k
Y
1 − zn = (1 − e2iπaj /nj z n/nj ),
j=1
as each n-th root of unity is a simple root of the polynomial on the left
hand side. This can be rewritten as
X P P
1 − zn = (−1)|I| z j∈S n/nj e2iπ j∈I aj /nj .
I⊂{1,··· ,k}
Here are two famous conjectures due to Erdös and others which are still
open.
Conjecture 1. (Erdös-Selfridge). There is no finite system of covering
congruences where all the moduli are distinct and odd.
Conjecture 2. (Erdös) For any M > 0, there exists a system of covering
congruences where all the moduli are distinct and > M .
Erdös mentioned in 1995 that the last conjecture is perhaps his favourite
problem!
We finish with a number-theoretic application which originally motivated
the study of covering congruences. This is the following beautiful result due
to (who else but!) Erdös.
Theorem. (Erdös, 1950). There are infinitely many odd integers which
are not expressible as the sum of a prime and a power of 2.
Proof. Consider a covering system of congruences ai (ni ) with 1 ≤ i ≤ k
where p1 , · · · , pk are distinct odd primes and ni the least positive number
satisfying 2ni ≡ 1 mod pi .
Erdös constructed such covering systems explicitly; one such is
101
Covering the Integers
Note that the ni ’s here – 2, 3, 4, 8, 12, 24 – are the orders of 2 modulo the
primes 3, 7, 5, 17, 13, 241.
Given any such system, the Chinese remainder theorem provides a com-
mon solution to the congruences
x ≡ 2ai (mod pi ); 1 ≤ i ≤ k and x ≡ 1Q(mod 2).
Such a solution x is unique modulo 2 ki=1 pi . Consider the smallest pos-
itive integer solution, say x0 . Now, for each integer r, there exists some i
such that r ≡ ai (mod mi ).
So, if n ≡ n0 (mod 2p1 · · · pk ), then
n − 2r ≡ n0 − 2ai ≡ 0 (mod pi ).
Thus, if n − 2r > pi , then it must be composite.
This proves that all n not of the form 2r + pi for some r and some i ≤ k
are inexpressible in the form asserted. To dispose of the exceptional cases,
we may impose extra congruence conditions.
An example (taking the above covering system 0(2), 0(3), 1(4), 3(8), 7(12),
23(24) of Erdös) gives us:
No integer ≡ 7629217 (mod 1184810) is a sum of a power of 2 and an
odd prime.
Z-W Sun has done substantial work in the subject of covering congru-
ences and revealed connections with zero-sum problems. We finish with
a following amazing application completed by Sun of the above work by
Erdös and later work of Cohen, Selfridge on covering congruences.
Consider the 29-digit number
M = 66483084961588510124010691590
= 2.3.5.7.11.13.17.19.31.37.41.61.73.97.109.151.241.257.331.
x ≡ 47867742232066880047611079 mod M
102
S Chowla and S S Pillai:
The Story of Two Peerless
Indian Mathematicians1
The chapter is a modified version of an article that first appeared in Resonance, Vol. 17, No. 9,
pp. 855–883, September 2012.
1 This chapter was part of an article written in collaboration with R Thangadurai.
103
The Story of Two Peerless Indian Mathematicians
“in figurative terms, what has been solved can be likened to an egg-shell, and
what remains to be solved to the infinite space surrounding it.”
104
The Story of Two Peerless Indian Mathematicians
105
The Story of Two Peerless Indian Mathematicians
106
The Story of Two Peerless Indian Mathematicians
107
The Story of Two Peerless Indian Mathematicians
former’s first visit there. After discussing mathematics, Chowla got them
both bottles of ‘pepsi’ from a vending machine. After the meeting,
Ramachandra says that he ran around the premises muttering that he drank
pepsi with Chowla!
Chowla’s fertile imagination earned him the sobriquet of ‘poet of mathe-
matics’ from his associates. Chowla passed away in the U.S. in 1995 at the
age of 88. On the other hand, Pillai died tragically at the age of 49 in 1950.
Pillai was invited to visit the Institute for Advanced Study in Princeton
for a year. The flight which he boarded to participate in the 1950 Inter-
national Congress of Mathematicians tragically crashed near Cairo on the
31st of August. The readers are directed to Pillai’s collected works which
have appeared recently.
is a sum of 2k + [(3/2)k ] − 2 numbers which are k-th powers and is not the
sum of a smaller number of k-th powers. Hence,
g(k) ≥ 2k + [(3/2)k ] − 2.
This ideal Waring conjecture asserts that this lower bound is the cor-
rect bound also. Pillai proved, among other things, that this ideal Waring
conjecture holds good under the condition on k that the remainder r on
dividing 3k by 2k satisfies r ≤ 2k − [(3/2)k ] − 2 (chapter 21 of [3]). This
is known to hold for all k ≤ 471600000. At this time, the ideal Waring
conjecture is known to hold for all large enough k.
108
The Story of Two Peerless Indian Mathematicians
and by Nn , the numerator of Bn /n, the numbers N20 , N37 are composite.
In 1930, Chowla showed [5] that Ramanujan’s claim has infinitely many
counter-examples. Surprisingly, Chowla returns to this problem 56 years
later(!) in a joint paper with his daughter [6] and poses as an unsolved
problem that Nn is always square-free. In a recent article, Dinesh Thakur
[7] points out that Chowla’s question has infinitely many negative
answers. Indeed, Thakur shows:
For any fixed irregular prime p less than 163 million, and any arbi-
trarily large k, there exists a positive integer n such that Nn is divisible
by pk .
109
The Story of Two Peerless Indian Mathematicians
If we observe (from the existing tables) that 372 divides N284 , Chowla’s
question has a negative answer. The proof of the more general assertion uses
the so-called Kummer congruences which essentially assert that the value of
(pn−1 −1)Bn
n modulo pk depends (for even n) only on n modulo pk−1 (p − 1),
if p − 1 does not divide n. Using this as well as certain functions called
p-adic L-functions, the general assertion of arbitrarily large powers can
also be obtained.
110
The Story of Two Peerless Indian Mathematicians
..
.
a1 x p−1 + a2 x1 + · · · + a p−1 x p−3 = 0.
2 2 2
If the ai ’s are not all 0, this leads to the vanishing of the ‘cirulant’
determinant
x1 x2 · · · x(p−1)/2
x2 x3 · · · x1
.
.. ..
. .
x(p−1)/2 x1 · · · x(p−3)/2
At this point, Chowla quotes the well-known value of this determinant
and proceeds.
What is the value of this determinant?
In the 3 × 3 case
x1 x2 x3
x2 x3 x1 ,
x3 x1 x2
has determinant 3x1 x2 x3 − x31 − x32 − x33 .
Let 1, ω, ω 2 be the cube roots of unity.
If x1 , x2 , x3 are replaced by ωx1 , ω 2 x2 , x3 or by ω 2 x1 , ωx2 , x3 , the expres-
sion remains the same. As x3 = −x1 −x2 leads to 3x1 x2 x3 −x31 −x32 −x33 = 0,
111
The Story of Two Peerless Indian Mathematicians
Here, Chowla realizes with surprise that the above factors are (up to
certain non-zero factors) none else other than the special value at s = 1
of certain functions L(s, χ) = ∞ χ(n)
P
n=1 n called Dirichlet L-functions cor-
responding to Dirichlet characters χ modulo p which satisfy χ(−1) = −1.
The non-vanishing of these are the older result mentioned above and show,
thus, that the determinant is non-zero. Hence, the linear independence of
the cot(rπ/p) for 1 ≤ r ≤ (p − 1)/2 is established.
Remarks. (i) Kai Wang closely followed Chowla’s proof to generalize his
theorem to non-primes and to derivatives of the cotangent function [11] by
showing:
φ(k)
s ≥0 and an arbitrary natural number k, the 2 real numbers
For any
ds rπ
dxs cot x+ k (where r ≤ φ(k)
2 and (r, k) = 1) are linearly independent
x=0
over Q.
(ii) The non-vanishing at s = 1 of L(s, χ) for nontrivial Dirichlet char-
acters is the key fact used in the proof of Dirichlet’s famous theorem on
existence of infinitely many primes in any arithmetic progression an + b
with (a, b) = 1.
(iii) Kenkichi Iwasawa (see [12]) showed later in 1975 that the above
result has connections with the so-called “regular” primes. The definition
of regular primes is not needed here and, we merely recall that Kummer had
proved that the Fermat equation xp + y p = z p has no non-zero solutions
in integers if p is a ‘regular prime’. It is unknown yet whether there are
infinitely many regular primes (although Fermat’s last theorem has been
proved completely) – surprisingly, it has been known for a long time that
there are infinitely many irregular primes! The connection of the above
result of Chowla with regular primes is the following.
112
The Story of Two Peerless Indian Mathematicians
p−1
2
2π X
2 sin = tr cot(2rπ/p).
p
r=1
Iwasawa shows the prime p is regular if and only if, none of the tr ’s have
denominators which is a multiple of p and, at least one tr has numerator
also not divisible by p.
113
The Story of Two Peerless Indian Mathematicians
This formula is useful in a number of ways. For instance, one can get
an asymptotic estimate of how fast An (d) grows with n (for any fixed d).
Moreover, for d = p, a prime, this gives a simple-looking closed formula for
An (p) for any n.
9.1 Closed Form for the Prime Case
In the above identity
X An (d) X xk
= exp .
n! k
n≥0 k|d
114
The Story of Two Peerless Indian Mathematicians
115
The Story of Two Peerless Indian Mathematicians
when d < 0 and a > 0, f (x, y) takes only positive values (excepting the
value 0 at x = y = 0). Thus, when a > 0, and the discriminant is negative,
the form is positive-definite.
For example, x2 + ny 2 is a positive-definite form with discriminant −4n.
Two forms f (x, y) and g(x, y) are said to be equivalent
or, said
to be in the
α β
same class if f (αx+βy, γx+δy)) = g(x, y) where A = ∈ SL(2, Z),
γ δ
an integer matrix of determinant 1.
The motivation behind this definition is:
Equivalent forms take the same sets of values as x, y vary over integers.
This is clear because one may also write
in the form
f (x, y) = g(δx − βy, −γx + αy)).
δ −β
Notice that the matrix is simply the inverse of the matrix
−γ α
α β
.
γ δ
Moreover, equivalent forms have the same discriminant. Gauss showed:
For d < 0, the number h(d) of classes of primitive, positive-definite binary
quadratic forms of discriminant d is finite.
Gauss conjectured that h(d) → ∞ as −d → ∞. This was proved by
Heilbronn. By a modification of Heilbronn’s argument, Chowla proved the
following fact which was another conjecture of Gauss [16]:
h(d)
2t → ∞ as −d → ∞ where t is the number of primes dividing d.
This interesting fact is useful in a totally different context which we
indicate briefly now.
Euler obtained the following list of 65 numbers called ‘Numerus idoneus’
(or ‘convenient’ numbers):
1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 12, 13, 15, 16, 18, 21,
22, 24, 25, 28, 30, 33, 37, 40, 42, 45, 48, 57, 58,
60, 70, 72, 78, 85, 88, 93, 102, 105, 112, 120, 130,
133, 165, 168, 177, 190, 210, 232, 240, 253, 273, 280,
312, 330, 345, 357, 385, 408, 462, 520, 760, 840, 1320, 1365, 1848.
Consider any odd number m co-prime to n which is expressible as
m = x2 + ny 2 with (x, ny) = 1. If each such m which has a unique ex-
pression of the form x2 + ny 2 in positive integers x, y is necessarily prime,
the number n is said to be idonean.
116
The Story of Two Peerless Indian Mathematicians
117
The Story of Two Peerless Indian Mathematicians
which shows that the corresponding matrices are conjugate in SL2 (Z).
The above associations are inverse to each other and proves the proposi-
tion.
Remarks. The above association is also useful in deciding if two
matri-
1 3
ces are conjugate in SL2 (Z) or not. For instance, the matrices
3 10,
1 1
which have trace 11 are associated to the quadratic forms
9 10
3x2 + 9xy − 3y 2 , x2 + 9xy − 9y 2 ,
respectively. However, they are evidently in-equivalent because the first one
takes only multiples of 3 as values whereas the second one takes values like
1 at (x, y) = (1, 0).
118
The Story of Two Peerless Indian Mathematicians
where the product varies over all the distinct prime divisors. This formula
shows that the functional value fluctuates a lot.
In analytic number theory, to study such fluctuating arithmetical func-
tions, one often looks at their average behaviour. One can prove that the
average value from 1 to x is π32 x but the interesting part is to have an idea
of the error which would be introduced if we take this value. In analytic
number theory, this methodology of ‘determining the main term and esti-
mating the error term’ is fundamental because we cannot deduce anything
concrete if the error term is of the same order as the main term! One has
X 3
φ(n) = 2 x2 + E(x),
π
1≤n≤x
119
The Story of Two Peerless Indian Mathematicians
120
The Story of Two Peerless Indian Mathematicians
and he produced infinitely many integers N for which the above bound is
attained.
For any given > 0, we can deduce from the above upper bound that
there is a constant N0 depending on so that
121
The Story of Two Peerless Indian Mathematicians
(where the product runs over all the prime numbers) converges absolutely
for Re(s) > 1 for all r ≥ 1. In particular, when s = 2, this series converges.
So,
∞
X d(n)r
< ∞ for all integers r ≥ 1.
n2
n=1
d(n)r
Therefore, the n-th term which is tends to zero. In particular, it
n2
is bounded for all large enough n’s. Thus, we get
for some M > 0 and c > 0 constants and this is true for all r ≥ 1.
Thus, for all n ≥ M , we get d(n) ≤ c1/r n2/r . Also, note that c1/r → 1 as
r → ∞.
Given > 0, we can find r0 such that 2/r < for all r ≥ r0 and we get
d(n) ≤ n .
To show how this sort of argument plays a role in the famous
Langlands conjectures, we describe such a conjecture. This can be done
through the famous Ramanujan delta function. Ramanujan studied the
following q-series
∞
Y
∆(z) = q (1 − q n )24 where z ∈ C with =(z) > 0; q = e2πiz
n=1
122
The Story of Two Peerless Indian Mathematicians
1 1
Since ∈ SL2 (Z) and by the above relation, we see that ∆(z+1) =
0 1
∆(z) and hence ∆ function is a periodic function.
Therefore, it has a Fourier expansion. It can be proved that the Fourier
expansion of ∆(z) is
X∞
∆(z) = τ (n)e2iπnz ,
n=1
where τ (n) is the Fourier coefficients which are integers.
Traditionally, one writes q = e2iπz so that ∆(z) = ∞ n
P
n=1 τ (n)q .
Ramanujan computed the initial τ values and conjectured the following
relations.
1. τ (mn) = τ (m)τ (n) whenever (m, n) = 1.
2. τ (pa+1 ) = τ (p)τ (pa ) − p11 τ (pa−1 ) for all primes p and a ≥ 1.
3. |τ (p)| < 2p11/2 for every prime p.
The first two conjectures were proved by Mordell in 1917 and the third one
was proved by P Deligne in 1975 using deep algebraic geometry.
Note that the third conjecture of Ramanujan is equivalent to the
assertion:
11
τ (n) = O(n 2 + )
for any given > 0.
Our interest is in this version of Ramanujan’s conjecture.
Let us define for each integer n ≥ 1,
τn = τ (n)/n11/2 .
Then Ramanujan’s conjecture is equivalent to τn = O(n ) for any given
> 0. Define the L-series attached to ∆ function as
∞
X τn
L(s, ∆) = ,
ns
n=1
where s ∈ C with =(s) > 0.
Since τ (n) is a multiplicative function (the first conjecture of Ramanujan
mentioned above and proved by Mordell), we see that τnr is also a multi-
plicative function and hence we get
Y τp τp2
L(s, ∆) = 1 + s + 2s + · · · .
p
p p
Using the second conjecture of Ramanujan, one notes that for all primes
p, we have
∞
X 1 1
τpa X a = 2
= ,
1 − τp X + X (1 − αp X)(1 − βp X)
a=0
123
The Story of Two Peerless Indian Mathematicians
References
[1] B Sury, Box 1, Resonance, Vol. 9, No. 6, 2004.
[2] Waring’s problem and the circle method, Resonance, Vol. 9, No. 6,
pp.51–55, 2004.
124
The Story of Two Peerless Indian Mathematicians
125
Multi-variable Chinese
Remainder Theorem
1. Introduction
The Chinese remainder theorem (CRT) seems to have originated in the
3rd century AD in the work of Sun-Tsu. There are also versions in Indian
5th century mathematics of Aryabhata. The classical versions dealt with
coprime moduli. Oystein Ore proved a version [1] for non-coprime moduli
in 1952 in the American Mathematical Monthly but, this does not seem
to be well-known because a paper published 50 years later by Howard [2]
proves the same result! However, a multi-variable version does not seem
to be known. We present such a version and point out that there are still
many questions open for investigation.
2. Classical CRT
A variant of a folklore tale goes as follows. Three thieves steal a number
of gold coins and go to sleep after burying the loot. During the night, one
thief wakes up and digs up the coins and, after distributing into 6 equal
piles, finds 1 coin left over which he pockets quietly after burying the rest
of the coins. He goes back to sleep and after a while, a second thief wakes
up and digs up the coins. After making 5 equal piles, he again finds 1 coin
left over which he pockets and buries the rest and goes to sleep. The 3rd
thief wakes up and finds the rest of the coins make 7 equal piles excepting
a coin which he pockets. If the total number of coins they stole is not more
than 200, what is the exact number?
With a bit of hit and trial, one can find that 157 is a possible number.
The Chinese remainder theorem gives a systematic way of solving this in
general.
In the above problem, the sought-for natural number N is so that N − 1
is a multiple of 6, N − 2 is a multiple of 5 and N − 3 is a multiple of 7. This
means that N leaves remainders 1, 2, 3 on division by 6, 5, 7 respectively.
Let us consider two coprime natural numbers m1 , m2 and suppose, we are
looking for a natural number N which leaves remainders a1 , a2 on division
by m1 , m2 respectively. Then, the Euclidean division algorithm tells us that
The chapter is a modified version of an article that first appeared in Resonance, Vol. 20, No. 3,
pp. 206–216, March 2015.
127
Multi-variable Chinese Remainder Theorem
m1 k1 + m2 k2 = 1.
The number
N = a1 m2 k2 + a2 m1 k1
has the property that (N −a1 ) is a multiple of m1 and (N −a2 ) is a multiple
of m2 . We have not yet got a number as sought since N could be negative.
However, we may add a suitable multiple of m1 m2 to N and that will satisfy
the requirements.
More generally, suppose there are r natural numbers m1 , · · · , mr which
are pairwise coprime. We seek a natural number N leaving given remainders
a1 , · · · , ar on divisions by m1 , · · · , mr respectively. An appropriate gener-
alization of the above argument for two numbers is the following. If Mi
denotes the product of all the mj ’s excepting mi , then the GCD of mi
and Mi is 1 for each i = 1, · · · , r. As above, the Euclidean algorithm gives
integers ni , ki such that
ni Mi + ki mi = 1,
for each i = 1, · · · , r.
We have therefore, ai ni Mi − ai is a multiple of mi for each i ≤ r.
As mi divides Mj for each j 6= i, the integer
N = a1 n1 M1 + a2 n2 M2 + · · · + ar nr Mr ,
128
Multi-variable Chinese Remainder Theorem
N = a1 n1 M1 + a2 n2 M2 + · · · + ar nr Mr .
129
Multi-variable Chinese Remainder Theorem
130
Multi-variable Chinese Remainder Theorem
131
Multi-variable Chinese Remainder Theorem
if, for each i ≤ k, there is some j for which aij is coprime to mi , then
the necessary condition obviously holds.
(iv) In the classical case of one variable, there is a unique solution modulo
m1 m2 · · · mk . In the multivariable case, there is no natural unique-
ness assertion possible. The point is that homogeneous congruences in
more than one variable have many solutions. So, uniqueness can be
asked for only after specifying a box (more precisely, an n-dimensional
parallelotope) in which we seek solutions.
For example, both (1, 4) and (0, −1) are simultaneous solutions of
the congruences
x − y ≡ 1 (mod 2),
x + y ≡ 2 (mod 3).
(v) The Euclidean division algorithm is the principal reason behind these
classical versions of the Chinese remainder theorem. In particular, it
holds good over the polynomial ring in one variable over a field. If n
elements are coprime, then there is a linear combination which gives 1.
This is no longer if we consider, for instance, polynomials in two vari-
ables.
For example, in the polynomial ring C[X, Y ], consider the congru-
ences
t ≡ 0 (mod X),
t ≡ 1 (mod Y ).
Here, of course, by a congruence f (X, Y ) ≡ g(X, Y ) mod h(X, Y ),
we mean that there is a polynomial k(X, Y ) so that
132
Multi-variable Chinese Remainder Theorem
and sufficient criterion for these cases, we will get such a criterion for the
general case. In this section, we fix a prime p and moduli mi = pti . We
further consider only the special case of n congruences in n variables. That
is, let us look at an n×n integer matrix A = {aij } and at the corresponding
system of congruences
n
X
aij xj ≡ bi (mod pti ) ∀ i ≤ n.
j=1
133
Multi-variable Chinese Remainder Theorem
References
[1] Oystein Ore, The general Chinese remainder theorem, The American
Mathematical Monthly, Vol. 59, No. 6, 1952.
[2] Fredric T Howard, A Generalized Chinese Remainder Theorem,
The College Mathematics Journal, Vol. 33, No. 4, pp. 279–282,
September 2002.
134
Which Positive Integers are Interesting?
30031
We are exposed to a beautiful thought process in school when we learn
Euclid’s proof of the infinitude of primes. Recall that the argument pro-
ceeds by observing that once we have gotten hold of the first few prime
numbers, the number obtained by adding 1 to their product must have
a prime factor which must necessarily be larger than the previous ones.
As this large number leaves remainder 1 on division by any of these primes,
any of its prime factors is larger than the previous primes (and hence gives
a new prime). The ‘hope’ (if one may call it that) that this new num-
ber is itself a prime, leads quickly to disillusionment. The first example is
2 × 3 × 5 × 7 × 11 × 13 + 1 = 30031. I leave it to the reader to find the
prime factors of this number. The intriguing question as to whether we do
get primes infinitely often in this process is still open! The largest known
prime P for which the product of all the primes until P is 1 less than a
prime number is 42209.
The chapter is a modified version of an article that first appeared in Resonance, Vol. 20, No. 8,
pp. 680–698, August 2015.
135
Which Positive Integers are Interesting?
561
In cryptography, one of the recurring themes is the employment of the
so-called Fermat little theorem – If p is a prime number and a is an integer
which is not a multiple of p, the number ap−1 − 1 is a multiple of p.
The thought that this property might characterize all primes perishes
soon. There are composite positive integers n such that every positive inte-
ger a co-prime to n possesses the property that an−1 − 1 is a multiple of n.
Such numbers – now known as Carmichael numbers – have a characterizing
property. This is the property:
n is a Carmichael number if and only if it is square-free and each prime
divisor p of n satisfies p − 1 divides n − 1.
Look at ‘Prime ordeal’, Resonance, pp.866-881, September 2008, for a
proof.
The smallest Carmichael number is 561. Indeed, 561 = 3 × 11 × 17 and
560 = 16 × 5 × 7.
If a is relatively prime to 561, then a560 −1 has factors a2 −1, a10 −1, a16 −1
which are multiples of 3, 11, 17 respectively, by Fermat’s little theorem.
15
For an integer n > 1, look at all its divisors including 1 and n. Let s(n)
denote the sum of all digits of all the divisors.
For example, s(10) = 1 + 2 + 5 + (1 + 0) = 9.
Let us iterate this process, that is, look at s2 (n) = s(s(n)), s3 (n) =
s(s2 (n)) etc. In general, sk+1 (n) = s(sk (n)).
For instance
s2 (10) = s(9) = 1 + 3 + 9 = 13,
s3 (10) = s(13) = 1 + 1 + 3 = 5,
s4 (10) = s(5) = 1 + 5 = 6,
s5 (10) = s(6) = 1 + 2 + 3 + 6 = 12,
136
Which Positive Integers are Interesting?
15 again!
The number 15 has a claim to fame for another reason too. To motivate
it, we recall a few things. Lagrange proved that every positive integer is
expressible as a sum of four squares of integers. On the other hand, Gauss
proved that a natural number n is expressible as a sum of three squares of
integers if and only if it is NOT of the form 4k (8r +7). Indeed, Gauss was so
excited about this discovery which he noted in his mathematical diary as:
“EYPHKA! ∆ + ∆ + ∆ = n.”
137
Which Positive Integers are Interesting?
It is said that this was the single discovery that turned Gauss’s mind
into taking up mathematics as a career although he was a great philologist
as well. Fermat stated the result that a positive integer > 1 is a sum of
two squares of integers if and only if, in its prime decomposition, every
prime of the form 4k + 3 appears with an even power. Ramanujan wrote
down a list of 55 such ‘positive forms’ ax2 + by 2 + cz 2 + dw2 for positive
integers a, b, c, d which he claimed were the only ones of this form which
take ANY positive integer value as the variables take integer values. His
list was almost perfect – the one exception x2 + 2y 2 + 5z 2 + 5w2 takes all
values excepting the value 15(!)
The mathematician and puzzlist John Conway came up with the follow-
ing general observation which he proved along with his student William
Alan Schneeberger. Consider
n
X
q(x1 , x2 , · · · , xn ) = aij xi xj
i,j=1
in n variables which takes only strictly positive values for all real values of
the variables other than xi = 0 for all i (one calls it positive-definite),
where all aij ’s are integers and aij = aji for i 6= j. They proved the
remarkable theorem that if this function takes all the integer values from
1 to 15 when we consider integer values for the variables xi , then it takes
ALL integer values! Conway–Schneeberger’s proof was very involved and
the mathematician Manjul Bhargava who received a Fields medal in 2014
(not for this work though) came up with a much simpler proof of this result
and, what is more, vastly generalized the result. Thus, 15 is special for the
reason that:
Pn
If i,j=1 aij xi xj is positive-definite, and aij are integers such that
aij = aji , and if all integer values from 1 to 15 occur as values of the
form when evaluated at suitable integers x1 , x2 , · · · , xn then ALL positive
integers occur as values. Moreover, 15 is the smallest such number.
1729
Any list of interesting positive integers is likely to include the taxicab
number 1729. The story of Ramanujan coming up with the observation
that 1729 is the smallest positive integer which is the sum of two perfect
cubes in two different ways
103 + 93 = 1729 = 123 + 13 ,
is too well-documented to repeat here. However, what may not be so well-
known is that 1729 is also a Carmichael number! Indeed, 1729 = 7 × 13 × 19
and 1728 = 26 × 33 . If a is coprime to 1729, then a1728 − 1 has factors
138
Which Positive Integers are Interesting?
1806
This number has a very curious origin. It turns out to be the unique
solution to the following problem.
Q
Find all the even numbers n which satisfy n = p prime,(p−1)|n p.
Note that this means n includes ALL possible primes p for which p − 1
divides n. Thus, numbers like n = 2, 6 are ruled out.
Note that n must be square-free. One easily sees that 2, 3, 6, 7, 43 divide n.
Moreover, any such n must be divisible by these numbers (and perhaps
others). Because of the hypothesis that a prime p|n if, and only if, (p−1)|n,
if a new prime factor of n arises, it must be one more than a product of
smaller prime factors of n. However, the above numbers cannot give a new
prime because the numbers
2 × 43 + 1, 2 × 3 × 43 + 1, 2 × 7 × 43 + 1, 2 × 3 × 7 × 43 + 1
are all composite. Therefore, the unique answer to the problem is the num-
ber 2. × 3 × 7 × 43 = 1806.
The discovery/appearance of this number is due to Kellner and is in
the context of Bernoulli numbers – the numerator of the n-th Bernoulli
number is the above product. Thus, 1806 is the unique number n for which
the numerator of Bn equals n.
6174
This was a discovery by D Kaprekar in the 1940’s. Starting with any
4-digit number (other than those with identical digits), apply the following
transformation. Arrange the digits in the descending order, say a > b >
c > d. Subtract the number with the digits dcba from abcd to obtain a
4-digit number (even if it is a 3-digit number, it should be regarded as a
4-digit number with 0 in the beginning). This transformation produces after
finitely many iterations (at the most 7), the number 6174 which has come
to be known as the Kaprekar constant. Note that 6174 is invariant under
this transformation.
Before proving that every 4-digit number leads to 6174, we should first
look for such constants among 2-digit and 3-digit numbers.
It is immediately seen that any 2-digit number (other than those with
identical digits) leads to the cycle
09 → 81 → 63 → 27 → 45 → 09.
139
Which Positive Integers are Interesting?
The unique 3-digit Kaprekar constant is 495. So, 495 has at least as much
claim to fame as 6174 (!)
Indeed, if abc is a 3-digit number with a ≥ b ≥ c and a > c, and if we
write
abc − cba = pqr,
then
10 + c − a = r, 10 − 1 + b − b = q, a − 1 − c = p.
Thus, q = 9 and p + r = 9. Hence, we need to check only the numbers
10 + d − a = s, 10 − 1 + c − b = r, b − 1 − c = q, a − d = p.
for
(p, q) = (9, 1), (8, 2), (7, 3), (6, 4), (5, 5).
Each of these leads to 6174.
After this, we could go in two different directions – look at a general
number d of digits or/and a general base b in place of 10. We mention a
few results and leave it to the interested reader to investigate further.
For instance, if the base b = 2r, then the only 3-digit Kaprekar constant
in base b has the digits r − 1, 2r − 1 and r – the proof of this generalization
is the same as that of 495 in base 10.
It can be proved that there are no odd bases b admitting a 3-digit
Kaprekar constant.
As for 4-digit Kaprekar constants, there is one in base 5 which is
3032 – so, once again this has as much claim to be of interest as
6174 has!
140
Which Positive Integers are Interesting?
The other bases where 4-digit Kaprekar constants exist are of the form
b = 4k × 10. In this base, the 4-digit Kaprekar constant has digits
6 × 4k , 2(4k − 1) + 1, 8(4k − 1) + 7 and 4 × 4k .
This can be proved similarly to the case of base 10.
There is no 5-digit Kaprekar constant in base 10.
On the other hand, base 15 has the Kaprekar constant with the 5 digits
10, 4, 14, 9, 5.
There exist 5-digit Kaprekar constants in each base of the form
b = 6k + 3 ≥ 15; this is left to the interested reader to determine.
3435
This is sometimes known as the Ramachandra number. An eminent
number theorist K Ramachandra observed when he was in college that
his Professor’s car number 3435 has the property
3435 = 33 + 44 + 33 + 55 .
This is the only number > 1 with this property. However, this remains
just a curiosity and does not seem to unveil any serious mathematics.
1848
We will see that 1848 is the largest of 65 numbers written down by
Euler with a certain property. It is known that there could be at the most
two larger numbers with that property. It is easy to show that if an odd
number has a unique expression as a sum of two squares of positive integers
n = x2 + y 2 and, if x, y are coprime, then n must be a prime number. Euler
generalized this property in order to obtain a primality criterion. He defined
a positive integer m to be ‘Idoneal’ or ‘convenient’ (‘Idoneus Numerus’ in
Latin) if it satisfies the property:
If an odd positive integer n admits a unique expression n = x2 + my 2
with x, y > 0 and if, in addition, the GCD (x, my) = 1, then n must be
prime.
Euler wrote down a list of 65 convenient numbers (the smallest ‘inconve-
nient’ number is 11) based on a criterion he obtained. The largest in his list
is 1848. Until date, no bigger convenient number has been found. S Chowla
was the first to prove in 1934 that there are only finitely many convenient
numbers. This is based on deep methods (coming under the umbrella of
class field theory) outside the scope of our discussion. Later, in 1973, it
has been shown by Weinberger that Euler could have missed at most two
other convenient numbers. Indeed, assuming the truth of a deep unsolved
141
Which Positive Integers are Interesting?
8191
We know that every positive integer can be represented in binary form
(that is, in base 2) in terms of 0’s and 1’s. There is nothing sacrosanct
(mathematically) about base 2 and, one may represent numbers in any
base one wants to use. Notice that the number 31 has the base 2 expansion
(11111)2
and the base 5 expansion
(111)5 .
So, it is natural to ask which natural numbers have all their digits to be
equal to 1 with respect to two different bases > 1.
It was observed by Goormaghtigh nearly a century ago that 8191 has
this property;
(111)90 = (1111111111111)2 .
In usual decimal (base 10) notation, this number is 8191.
The question can be posed in another form as follows. If b1 6= b2 are
two positive integers > 1, then the number with m ones in base b1 is
2 m−1 bm
1 −1
1 + b1 + b1 + · · · + b1 = b1 −1 . Therefore, we are asking if this number
can consist of n one’s in another base b2 .
This is equivalent to solving
xm − 1 yn − 1
=
x−1 y−1
in natural numbers x, y > 1 for some m, n > 2.
The largest known solution is 8191 mentioned above. It is still unknown
whether there are only finitely many solutions in all variables x, y, m, n. In
fact,
1093
Fermat’s last ‘theorem’ – asserting that the equation xn +y n = z n has no
solutions in positive integers x, y, z when n > 2 – took 350 years to be justi-
fiably called a theorem. However, there were several subjective results from
the old times. One of them due to Wieferich showed that the first case of
Fermat’s last theorem holds good for a prime p for which 2p−1 − 1 is not a
multiple of p2 . That is, for such a prime p, the equation xp + y p = z p
142
Which Positive Integers are Interesting?
bp−1 6≡ 1 mod p2 .
143
Which Positive Integers are Interesting?
1 + bk + b2k + · · · + bnk
p ≥ bk + 1 > bk − 1.
p−1
bp−1 ≡ 1 + p(bk − 1) mod p2 .
(n + 1)k
144
Which Positive Integers are Interesting?
71
John Conway discovered an amazing fact. Start with any positive integer
other than 22. Let us start with 1 say. Define the sequence which just reads
out the number of times each chain of digits is repeated in turn. That is,
after 1, we have 11 (meaning one 1) and after that we have 21 (to mean two
1’s) and 1211 (to mean one 2, one 1) and 111221 (meaning one 1, one 2,
two 1’s) etc. In general, if ak11 ak22 · · · akr r with ai 6= ai+1 , then the next term
145
Which Positive Integers are Interesting?
Skewes’s constants
The great mathematician C F Gauss conjectured at the age of 15, what
is now called the prime number theorem. He conjectured that the number
π(x) of primes not exceeding a number R xx is asymptotically given by the
logarithmic integral function li(x) = 2 logdt t . Here, by ‘asymptotically’,
we mean that the ratio π(x)/li(x) approaches 1 as x grows unboundedly
large. However, the inequality π(x) < li(x) was seen to hold for values of
146
Which Positive Integers are Interesting?
Graham’s constant G
The number is so gigantic that additional notation is needed to write it
down. This number arose as follows.
Consider a hypercube in n dimensions. This is the generalization of a
square in 2 dimensions and a cube in 3 dimensions; it has 2n vertices. If we
join every vertex to every other one, we get what is known as a complete
graph. R L Graham and B L Rothschild considered the following problem.
If we colour each edge with one of two available colours, is it always true
that there must exist a complete subgraph containing four coplanar vertices
such that all its six edges are of the same colour?
This is not necessarily true for 3-dimensional cubes – we leave it to the
reader to construct an example. On the other hand, Graham and Rothschild
proved the existence of a complete, monochromatic subgraph containing
four coplanar vertices in any colouring, if the dimension n is large enough.
Until now, one does not know the minimal possible value of n with this
property but the proof of Graham and Rothschild showed the existence
of an n which is at the most a constant G known as Graham’s constant.
To define what G is, we introduce Knuth’s up-arrow notation.
For positive integers a, b we already know the usual exponentiation ab as
a shorthand notation for multiplying a to itself b times. Knuth introduces
the up-arrow notation as: a ↑ b for ab . Next, define
a ↑↑ b = a ↑ (a ↑ (a ↑ (· · · ))) .
| {z }
b times
4 (33 ) 27
For example, 4 ↑↑ 3 = 4(4 ) while 3 ↑↑ 4 = 3(3 ) = 3(3 ) a much larger
number. In fact, the former has about 154 digits whereas the latter has
147
Which Positive Integers are Interesting?
a ↑↑↑ b = a ↑↑ (a ↑↑ (a ↑↑ (· · · ))) .
| {z }
b times
More generally,
g1 = 3 ↑4 3, g2 = 3 ↑g1 3, · · · , gn = 3 ↑gn−1 3.
148
Counting, Recounting and Matching
1. Introduction
Although we start our mathematical education by learning to count, it
is this very thing which is one of the most difficult things to carry out in
practice. In this article, we discuss a very simple method which is available
in a very general form under the name of Fubini’s principle. In its simplest
form, it is the obvious observation that counting in two different ways
produces the same sum. However, the various examples we discuss will
hopefully convince the reader that the principle is surprisingly effective.
The chapter is a modified version of an article that first appeared in Resonance, Vol. 21, No. 4,
pp. 353–368, April 2016.
149
Counting, Recounting and Matching
3. Valuable Tiles
Tile the plane by unit squares with vertices at integer lattice points.
Inside each unit square, fill in some real number and call it the value of
that unit square. Let A be a finite collection of unit squares having the
property that the total value of the ‘translate’ A + (i, j) is positive for
each lattice point (i, j). Here, by translate A + (i, j), we mean the set
{(x + i, y + j) : (x, y) ∈ A}. Then, we claim that for each finite collection
B of unit squares, some translate of B must have positive value.
The solution depending on the Fubini principle goes as follows:
Denote by [i, j], the unit square whose lower left corner has co-ordinates
(i, j) and let v(i, j) denote its value. Write
A = {[i1 , j1 ], · · · , [ir , jr ]}
which gives
r X
X s
( v(in + km , jn + lm )) > 0.
n=1 m=1
Therefore, we must have some n ≤ r for which sm=1 v(in +km , jn +lm ) >
P
0; hence, the translate (in , jn ) + B has total positive value.
4. Partitions
A well-known application of the Fubini principle is the following stunning
fact about partitions of a number. Recall that the number p(n) of partitions
of n is the number of ways of partitioning n objects into smaller collections.
For instrance p(4) = 5 because there are 5 ways to partition 4 objects:
150
Counting, Recounting and Matching
5. Regular Solids
Here is yet another striking example of the Fubini principle:
There are exactly five platonic solids.
In a platonic solid, the faces meeting at a vertex are regular polygons
with the same number of sides, the number of faces meeting at each vertex
is the same, and the solid angles at each vertex is the same.
If there are v vertices, e edges and f faces in the solid, and the faces
are regular polygons with p sides, then let us count the edges by counting
faces.
We get pf edges but we have counted each edge twice as it is the edge
of exactly two faces.
Hence, the Fubini principle implies 2e = pf .
Similarly, counting the edges by means of the two end point vertices, this
number also equals qv where q is the number of faces meeting at any vertex.
Hence,
qv = 2e = pf.
Rewrite it as
v e f
= = .
1/q 1/2 1/p
Now, we use the famous formula of Euler: v − e + f = 2; see, for instance
Example 9 on p.601 of the article ‘Invariants’ by B V Rajarama Bhat in
the July 2010 issue of Resonance. So,
v e f v−e+f
= = = .
1/q 1/2 1/p 1/q − 1/2 + 1/p
This is equal to
2 4pq 4pq
= = .
1/q − 1/2 + 1/p 2p − pq + 2q (4 − (p − 2)(q − 2)
151
Counting, Recounting and Matching
4pq
We had shown qv = 2e = pf = (4−(p−2)(q−2) .
Therefore, (p − 2)(q − 2) < 4 which gives exactly five solutions
(3, 3), (3, 4), (3, 5), (4, 3), (5, 3).
152
Counting, Recounting and Matching
and each pair of students has exactly one common teacher. Then, can we
determine the total number of students?
Let us suppose t is the total number of teachers (in our case t = 50) and
that the total number of students is s. Suppose each teacher teaches a total
of exactly s0 students (we have s0 = 57) and that each pair of students has
exactly t0 common teachers (we have t0 = 1).
We will find the relations between t, t0 , s, s0 .
Let us look at any teacher T and a pair of students Si , Sj taught by
her. Count the number of triples (T, Si , Sj ) as T varies over teachers and
Si , Sj vary over pairs of students such that T teaches both Si and Sj . As a
s0
teacher teaches exactly s0 students, there are exactly 2 triples (T, Si , Sj )
containing any particular teacher T . Therefore, the total number of triples
s0
is t 2 .
On the other hand, for each pair of students Si , Sj there are exactly t0
common teachers which means there are t0 triples (T, Si , Sj ) containing the
s
pair Si , Sj . As there are 2 ways to select the pair of students Si , Sj , the
total number of triples is t0 2s .
153
Counting, Recounting and Matching
ber of triples is t 32 . On the other hand, for any fixed pair (S1 , S2 ) among
the 9 students, there is exactly one common supervisor so that the number
of triples is 92 . By Fubini’s principle,
3 9
t = .
2 2
154
Counting, Recounting and Matching
∅=
6 S1 ⊂ S2 ⊂ · · · ⊂ Sn = S,
155
Counting, Recounting and Matching
In other words, Pm 1 n ≤ 1.
j=1 (kj )
n
where it is understood that a binomial coefficient d = 0 if n < d.
Note the particular case
X g 2 2g
= .
u g
u≥0
156
Counting, Recounting and Matching
E 7→ ex ex , W 7→ wx wx , N 7→ ex wx , S 7→ wx ex
is a bijection.
Hence, we obtain
X
4n + 2 2n+1−2r 2n + 1 2r
= 2 .
2n + 1 r
2r r
157
Counting, Recounting and Matching
Indeed, the first term 2tn counts the number of tilings of the 1 × (n + 1)-
board with the last tile a square, and tn−1 counts those for which the last
tile is a domino.
We make a divisibility observation
Pn now:
tn divides t2n+1 as well as r=0 t2r .
In fact, look at the two possibilities for a tiling – one in which the n-th
and (n + 1)-th places are filled by a domino or not.
If they are not occupied by a domino, the board is “breakable” at the
n-th place, and there are tn tn+1 such tilings.
If a domino occupies the n-th and (n + 1)-th squares, the number of
tilings is tn−1 tn .
Hence, we have
t2n+1 = tn tn+1 + tn−1 tn = tn (tn+1 + tn−1 )
which is evidently a multiple of tn .
Note also that since tn+1 , tn−1 have the same
P parity, tn+1 /2 is a multiple
of tn which means that tn divides t2n+1 /2 = r≥0 t2r .
On the other hand, in any tiling of the 1 × n board, consider the number
d of dominos.
They occupy 2d squares, and in the rest of the n − 2d squares, one can
have a black or white square which gives 2n−2d possibilities.
Now, the number of tiles here is n − d (because there are d dominos and
n − 2d unit squares).
number of ways to choose d dominos from these n − d tiles
So, the
n−d
is d .
Hence, we get:
n−2d n − d
X
tn = 2 .
d
d≥0
4n
X
(t2n−1 + t2n )2 = tr .
r=0
I also leave it to the interested person to discover such identities when
we count tilings by squares of a different colours and dominos of b different
colours.
158
Counting, Recounting and Matching
(2, 1, 0), (3, 1, 0), (3, 2, 0), (3, 2, 1), (4, 1, 0), (4, 2, 0), (4, 2, 1), (4, 3, 0), · · · .
Then, we have:
Given any positive integer n, and any base b such that rb > n, the (n+1)-
159
Counting, Recounting and Matching
expansion and can simply be proved by the greedy algorithm but the above
statement gives a combinatorial interpretation.
Here are a couple of examples to illustrate the statement.
(i) Let r = 3 and n = 12.
We may take any base b so that 3b > 12. For example, b = 6 is allowed
because 63 = 20.
References
[1] B Sury, Macaulay expansion, Amer. Math. Monthly, Vol. 121, 2014.
160
Odd if it isn’t an Even Fit!
Lighting up Tiling
There is some interesting history behind the discovery of the above fact.
In 1965, Fred Richman from the university of New Mexico had decided to
pose this in an examination in the master’s programme. He had observed
this in some cases but when he tried to prove it in general prior to posing
it in the exam, he was unsuccessful. So, the problem was not posed in the
exam. His colleague and bridge partner John Thomas tried for a long time
and finally came up with a proof that it is impossible to break the unit
square with corners at (0, 0), (1, 0), (0, 1)(1, 1) cannot be broken into an
odd number of triangles when the vertices of all the triangles have rational
co-ordinates with odd denominators. He sent the paper to the Mathematics
Magazine where the referee thought the result may be fairly easy (but could
The chapter is a modified version of an article that first appeared in Resonance, Vol. 20, No. 1,
pp. 23–33, January 2015.
161
Lighting up Tiling
not find a proof himself) and perhaps known (but could not find a refer-
ence to it). On the referee’s suggestion, Richman and Thomas posed this
as a problem [1] in the American Mathematical Monthly which nobody
could solve. Subsequently, Thomas’s paper appeared in the Mathematics
Magazine [2] in 1968. Finally, in 1970, Paul Monsky proved the complete
version in a paper [3] in the American Mathematical Monthly removing the
restriction imposed in Thomas’s paper.
We shall discuss Monsky’s proof of this beautiful fact presently.
Amazingly, it uses some nontrivial mathematical objects called 2-adic
valuations.
In fact, one may consider generalizations of squares like cubes and hyper-
cubes of higher dimensions. If an n-dimensional cube is cut into simplices
(generalizations of triangles like tetrahedra etc. in higher dimensions) of
equal volumes, it turns out that the number of simplices must be a multi-
ple of n!
One also says that a region like the interior of a square is ‘tiled’ by
triangles if the square can be broken into triangular pieces.
There is also a generalization of the result on tiling by triangles of
squares to the (so-called) polyominos. Polyominos are just unions of unit
squares.
Connected subsets of the square lattice tiling of the plane are called
special polyominos. That is, they have standard edges – edges of the unit
squares are parallel to the co-ordinate axes. The generalization alluded to
is due to S K Stein [4] and asserts:
Consider a special polyomino which is the union of an odd number of
unit squares. If this polyomino is a union of triangles of equal areas, then
the number of triangles is even.
We discuss the proof of this statement after discussing the solution by
Paul Monsky of the first problem we started with. It can be noticed that
the proof uses crucially that the number of unit squares in the polyomino
is odd. Interestingly, this question is still unanswered when the number of
unit squares in the polyomino is even.
The proof of Monsky as well as Stein’s result above use the so-called
‘2-adic valuation function’. The 2-adic valuation is a function from the set
of non-zero rational numbers to the set of integers; it simply counts the
power of 2 dividing any integer or, more generally, any rational number.
What is meant by the power of 2 dividing a rational number? Writing any
non-zero rational number as p/q with p, q having no common factors, one
may look at the power of 2 dividing p or q. If p, q are both odd, this is
simply 0. If p is even, then the 2-adic valuation of p/q is defined to be the
power of 2 dividing p. If q is even, then the 2-adic valuation of p/q is defined
to be the negative of the power of 2 dividing q, Formally, we write:
162
Lighting up Tiling
φ : Q∗ → Z
a
defined by φ( 2c b ) = a where b, c are odd; define also φ(0) = ∞ so that φ is
defined for all rational numbers. We keep in mind that 0 has larger 2-adic
value than any other rational number.
Colour a point (x, y) ∈ Q × Q by the colour:
φ:R→R
which satisfies:
φ restricts to the 2-adic valuation on Q;
φ(xy) = φ(x) + φ(y);
φ(x + y) ≥ min(φ(x), φ(y)).
This is important to have such an extension because we would really like
to colour all points in a square.
For instance, as φ(3/4) = φ(2−2 3) = −2, the second property above
implies that √
φ(3/4) = 2φ( 3/2) = −2.
√
Hence φ( 3/2) = −1.
Let us start Monsky’s proof by first considering the unit square with a
left lower corner at (0, 0) and making a few easy observations:
(i) If a point a is red, then any point x and x + a have the same colour.
(ii) On any line, there are at the most two colours.
(iii) The boundary of the square has an odd number of segments which
have a red end and a blue end.
(iv) If a triangle is not ‘complete’ (that is, has vertices only of one or two
colours), then it has 0 or 2 red-blue edges.
163
Lighting up Tiling
Proof. Recall
164
Lighting up Tiling
165
Lighting up Tiling
We note that a standard (unit) edge with a blue end and a green
end must be parallel to the X-axis and lies on a line whose height
is odd.
Therefore, on the border of each standard square, there is an edge with
a blue end and a green end.
Edges in the interior of the polyomino are adjacent to two standard
squares whereas those on the boundary are adjacent to one standard square
of the polyomino.
As there is an odd number of standard squares, the above observation
applies and, implies that φ(2A) ≤ φ(n). But, φ(2A) ≥ 1; so n must be a
multiple of 2.
166
Lighting up Tiling
a a+1
1 ζ ζ 2 · · · · · · 1/ζ 1 ζ ···
ζ ζ2 ζ3 · · ·
ζ2 ζ3 · · ·
..
.
Since each tile (copy of the smaller rectangle used) contains all the a-th
roots of unity exactly once, and as the sum 1 + ζ + · · · + · · · ζ a−1 = 0, the
sum of all theP entries
Pnof the m × n rectangle must be 0.
Therefore, m i=1 j=1 ζ i+j−2 = 0.
167
Lighting up Tiling
References
[1] F Richman and J Thomas, Problem 5471, The American Mathematical
Monthly, 74, p. 329, 1967.
[2] J Thomas, A dissection problem, The Mathematics Magazine, 41,
pp. 187–190, 1968.
[3] P Monsky, On dividing a square into triangles, The American Mathe-
matical Monthly, 77, pp. 161–164, 1970.
[4] S K Stein and S Szabo, Algebra & Tiling, The Carus Mathematical
Monographs 25, Published by the Mathematical Association of America,
1994.
168
Polya’s One Theorem with 100 pages
of Applications
The chapter is a modified version of an article that first appeared in Resonance, Vol. 19, No. 4,
pp. 338–346, April 2014.
169
Polya’s One Theorem with 100 pages of Applications
same shape can be placed in more than one receptacle. Each shape is given
a value and the value of a configuration is the total value of all the shapes.
A typical problem would be to determine the number of configurations with
a given value. Those permutations of the receptacles which yield another
configuration which is equivalent to the original one, give rise to a group
and the theory is stated in terms of this group. For instance, in the cube-
colouring problem, let D be the set of 6 faces of the cube and R is the set
of two colours black and white. A configuration here is a colouring of the
faces by the two colours; that is, a mapping φ : D → R.
The group of rotations of the 6 faces has 24 elements (see the figure):
(i) the identity element;
(ii) rotating clockwise by 90 degree about the line through the centers of
opposite faces like A and D (there are three such rotations);
(iii) rotating by 180 degrees about the line through the centers of opposite
faces like A and D (there are three such);
(iv) rotating anti-clockwise by 90 degrees about the line through the cen-
ters of opposite faces like A and D (there are three such);
(v) rotating by 180 degrees about lines through the centers of diagonally
opposite edges like the edge of the faces A and B and the edge of the
faces D and E (there are six such);
(vi) rotating clockwise by 120 degrees about the line connecting diagonally
opposite vertices like the vertex where A,B,C meet and the vertex
where D,E,F meet (there are four such);
(vii) rotating anti-clockwise by 120 degrees about the line connecting
diagonally opposite vertices like the vertex where A,B,C meet and
the vertex where D,E,F meet (there are four of these).
A B
170
Polya’s One Theorem with 100 pages of Applications
26 + 6 · 23 + 8 · 22 + 3 · 22 · 22 + 6 · 22 · 2 = 240.
Therefore, Cauchy–Frobenius–Burnside lemma gives the number of orbits
to be 240/24 = 10.
To describe Polya’s theorems, we shall consider only finite sets D, R like
in the example of the cube. For a group G of permutations on a set of n
elements and variables s1 , s2 , . . . , sn , one defines a polynomial expression
(called the cycle index) for each g ∈ G. If g ∈ G, let λi (g) denote the number
of i-cycles in the disjoint cycle decomposition of g. Then, the cycle index of
G, denoted by z(G; s1 , s2 , . . . , sn ) is defined as the polynomial expression
1 X λ1 (g) λ2 (g)
z(G; s1 , s2 , . . . , sn ) = s1 s2 . . . snλn (g) .
|G|
g∈G
For instance,
X sλ1 1 sλ2 2 . . . sλk k
z(Sn ; s1 , s2 , . . . , sn ) =
1λ1 λ1 !2λ2 λ2 ! . . . k λk λk !
λ1 +2λ2 +...+kλk =n
171
Polya’s One Theorem with 100 pages of Applications
There are 10 distinct coloured cubes in all, using two colours, as we saw
above.
Incidentally, if we look at a regular octahedron, then the group of sym-
metries of the 6 vertices is the same group above of symmetries of the 6
faces of the cube.
c(x) = c0 + c1 x + c2 x2 + c3 x3 + · · ·
a0 + a1 x + a2 x2 + a3 x3 + · · ·
172
Polya’s One Theorem with 100 pages of Applications
For simplicity, suppose R has two elements Black and white which have
values 0 and 1. Then, the generating polynomial above is simply 1 + x.
Let us discuss the example of chlorination of Benzene where some hydro-
gen atoms get substituted by Chlorine atoms. Give the values 0 and 1 to Cl
and H, and note that the group of symmetries of the Benzene molecule is
the group of rotations of the regular hexagon, which is the so-called dihedral
group D6 of order 12. The cycle index of D6 is
1 6 3 2 2 2
z(D6 ) = s + 4s2 + 2s3 + 3s1 s2 + 2s6 .
12 1
Substituting 1 + xr for sr ’s, we obtain the polynomial
1 6 2 3 3 2 2 2 2 6
(1 + x) + 4(1 + x ) + 2(1 + x ) + 3(1 + x) (1 + x ) + 2(1 + x )
12
= 1 + x + 3x2 + 3x3 + 3x4 + x5 + x6 .
Therefore, the number of configurations which have 2 chlorine atoms
is the coefficient of 4 which is 3. These are the ortho dichlorobenzenes,
meta dichlorobenzenes and para dichlorobenzenes where the gap between
the vertices corresponding to the carbon atoms to which the two chlorine
atoms are attached are 1, 2 and 3 edges respectively.
More general weighted versions of Polya’s theorem as well as the immedi-
ate applications to enumerating isomers of chemical compounds have been
discussed in detail in the earlier article in 2002.
Graph Enumeration
A key application of Polya’s theorem is to the enumeration of graphs.
Indeed, the introduction of Polya’s paper begins with the words (as trans-
lated by Read):
“This paper presents a continuation of work done by Cayley. Cayley has
repeatedly investigated combinatorial problems regarding the determination
of the number of certain trees. Some of his problems lend themselves to
chemical interpretation: the number of trees in question is equal to the num-
ber of certain theoretically possible chemical compounds.”
Indeed, a chemical compound with no multiple bonds corresponds to a
tree where different types of vertices correspond to different atoms. In case
of multiple bonds, one may regard different kinds of edges also.
A tree consists of vertices and edges and is a connected graph where each
edge connects two vertices. There can be several edges meeting at a vertex.
There is no closed path. Therefore, the number of edges is one less than the
number of vertices. A vertex is called r-edged if there are exactly r edges
originating there.
173
Polya’s One Theorem with 100 pages of Applications
Consider an alkane – this has a formula Cn H2n+2 . The carbon atoms are
usually assumed to have valency 4 which means that the structure of the
alkane is determined (that is the positions of the hydrogen atoms are uniquely
determined) by the structure formed by the carbon atoms. Topologically dif-
ferent trees with n four-edged vertices and 2n + 2 one-edged vertices corre-
spond to the different isomers with the molecular formula Cn H2n+2 .
Thus, the enumeration of isomers as above is equivalent to the enumer-
ation of trees as above. Interestingly, in Polya’s paper, he describes the
groups of symmetries for certain chemical compounds as so-called wreath
products. Polya calls wreath products as coronas. As far as one can ascer-
tain, this is the first introduction and study of finite wreath products as
permutation groups.
Polya’s theorem was generalized by de Bruijn in a way which allows one
to permute the shapes in R also.
Musings on Music
Polya’s theorem has been applied to the theory of music. One may deter-
mine the number of chords. To define this, one takes the n-scale to be the
integers from 0 to n − 1 under addition modulo n. There are translations
a 7→ a + i, where 0 ≤ i < n. An equivalence class (that is an orbit) is
called a chord and, one wishes to determine for each r < n, the number
of r-chords; that is, the number of orbits consisting of r elements. This is
equivalent to colouring the n-notes by two colours – we choose the notes
in the chord by colouring them by one colour and those which are not in it
by the other colour. The group is simply the cyclic group of order n whose
cycle index is:
1X n/d
z(Cn ; s1 , · · · , sn ) = φ(d)sd .
n
d|n
174
Polya’s One Theorem with 100 pages of Applications
if n is even and
1 X n d 1 n−1
z(Dn ; s1 , · · · , sn ) = φ( )sn/d + s1 s2 2
2n d 2
d|n
if n is odd;
1 X n/d 1 n/2
φ(d) +
2n r/d 2 r/2
d|(n,r)
1 n2 − 1
1 X n/d
φ(d) +
2n r/d 2 [r/2]
d|(n,r)
References
[1] G Polya, Kombinatorische Anzahlbestimmungen fur Gruppen, Graphen
und chemische Verbindungen, Acta Mathematica, Vol. 68, pp. 145–254,
1937.
[2] J H Redfield, The theory of group reduced distributions, Amer. J.Math.,
49, pp. 433–455, 1927.
[3] Shriya Anand, How to count – an exposition of Polya’s theory of enu-
meration, Resonance, pp. 19–35, September 2002.
[4] David L Reiner, Enumeration in music theory, The American Mathe-
matical Monthly, pp. 51–54, January 1985.
[5] G Polya and R C Read, Combinatorial enumeration of groups, graphs,
and chemical compounds, Springer-Verlag, 1987.
[6] V Krishnamurthy, Combinatorics – Theory and Applications, Affiliated
East-West Press Private Limited, 1985.
175