Sunteți pe pagina 1din 89

CONTENTS

CHAPETER I INTRODUCTION AND PRELIMINARIES


GROUPS
1.1 Introduction
1.2 Preliminaries

CHAPTER II MÖBIUS INVERSION THEOREM OVER A LATTICE


OF SUB
2.1 Definitions and some properties of group actions
2.2 Möbius inversion theorem over a lattice of subgroups

CHAPTER III SOME CONGRUENCES ON BINOMIAL


COEFFICIENTS
AND STIRLING NUMBERS OF BOTH KINDS IN
THE CASE
G ═ Cp
3.1 Möbius function defined on L(CP )
3.2 Congruence relation on Binomial coefficients in the case G = C p
3.3 Congruence relations on Stirling numbers of second kind in the
case
G = Cp
3.4 Congruence relations on Stirling numbers of first kind in the case
G = Cp

CHAPTER IV SOME CONGRUENCES ON BINOMIAL


COEFFICIENTS
AND STIRLING NUMBERS OF BOTH KINDS IN
THE CASE
G ═ Cp × Cp
4.1 Möbius function defined on L ( C p × C p )
4.2 Congruence relation on Binomial coefficients in the case
G = Cp × Cp
4.3 Congruence relation on Stirling numbers of second kind in the
case
G = Cp × Cp
4.4 Congruence relation on Stirling numbers of first kind in the case
G = Cp × Cp

CHAPTER V SOME CONGRUENCES ON BINOMIAL


COEFFICIENTS
AND STIRLING NUMBERS OF BOTH KINDS IN
THE
GENERAL CASE G = C p
r

Möbius function defined on L ( C p )


r
5.1
5.2 Congruence relation on Binomial coefficients in the case G = C p
r

5.3 Congruence relation on Stirling numbers of second kind in the


case G = C p
r

5.4 Congruence relation on Stirling numbers of first kind in the case


G = C pr

CHAPTER I

INTRODUCTION AND PRELIMINARIES

1.1 INTRODUCTION

In 1872, J. Peterson[19] used the action of the cyclic group of


order p to prove the congruences of Fermat and Wilson. In 1980, Rota
and Sagan [24] investigated congruences derived from group acting on
functions, Gessel[11] has studied congruences for groups acting on
graphs and Bruce E. Sagan[26] has derived congruences for group
acting on a set of subsets, partitions or permutations. Here we have
made a little attempt to survey some of congruences for binomial
coefficients, Stirling numbers of second kind, Stirling numbers of first
kind and some other related congruences by using action of finite
abelian groups in different cases.
In this dissertation for notational convenience we let

[ n] = { 1, 2,..., n} and m + [ n ] = { m + 1, m + 2,..., m + n} . We take G to be the

abelian subgroup of the symmetric group which is a product of cyclic

groups Cm1 × Cm2 × ... × Cmr , where Cmr is generated by the cycle

( J + 1, J + 2,..., J + m ) , J = ∑ mi , and H ≤ G means H is a subgroup of G.


j
i< j

Finally the letter p will always be used to denote a prime number.

This work has been divided in to FIVE CHAPTERS.

Chapter I is devoted to fundamental definitions and properties of


Binomial coefficients, Stirling numbers of both kinds and some other
related definitions.

Chapter II is Möbius inversion theorem over lattice of subgroups.


It has two sections. In section 2.1, we present definitions, examples
and some properties action of groups. In section 2.2 we define Möbius
function on lattice of sub groups, we prove the Möbius inversion
theorem over lattice of subgroups and its corollary.

Chapter III is devoted some congruences on binomial

coefficients and Stirling numbers of both kinds in the case G = C p . It


consists of four sections. In section 3.1 we prove Möbius function

defined on L ( C p ) and corresponding theorem. In section 3.2, we prove

congruence relation on Binomial coefficients in the case G = C p . In


section 3.3, we prove congruence relation on Stirling numbers of

second kind in the case G = C p and its related theorems. In section 3.4,
we prove congruence relation on Stirling numbers of first kind in the

case G = C p and its related theorems.


Chapter IV is devoted some congruences on Binomial

coefficients and Stirling numbers of both kinds in the case G = C p × C p . It


consists of four sections. In section 4.1, we prove Möbius function

defined on L ( C p × C p ) and its corresponding theorem. In section 4.2, we

prove congruence relation on binomial coefficients in the case

G = C p × C p . In section 4.3, we prove congruence relation on Stirling

numbers of second kind in the case G = C p × C p and its related theorem.


In section 4.4, we prove congruence relation on Stirling numbers of

first kind in the case G = C p × C p .

Chapter V is devoted some congruences on Binomial


coefficients and Stirling numbers of both kinds in the general case

G = C pr . It consists of four sections. In section 5.1 we prove Möbius

function defined on L ( C p ) and its corresponding theorem. In section


r

5.2, we prove congruence relation on binomial coefficients in the case

G = C pr . In section 5.3, we prove congruence relation on Stirling

numbers of second kind in the case G = C p . In section 5.4, we prove


r

Congruence relation on Stirling numbers of first kind in the case G = C p .


r

1.2 PRELIMINARIES
In this section, the fundamental definitions and results which will
be used in our work are presented.
Definition 1.2.1[6]: The ‘falling factorial polynomial’ is denoted by

( x) n and is defined by

( x ) n = x ( x − 1) ( x − 2 ) …( x − n + 1)
In particular, ( x ) 0 = 1 , since it is an empty product.

Examples 1.2.2:

(1) ( x)1 = x

(2) ( x ) 2 = x ( x − 1) = x2 − 1

(3) ( x ) 3 = x ( x − 1) ( x − 2 ) = x3 − 3x2 − 2 x

(4) ( x ) 4 = x ( x − 1) ( x − 2 ) ( x − 3) = x 4 − 6 x3 + 7 x 2 + 6 x

Definition 1.2.3 [6]: The ‘rising factorial polynomial’ is denoted by


x ( n ) and is defined by
x ( n ) = x ( x + 1) ( x + 2 ) …( x + n − 1) .

In particular, x (0) = 1 , since it is an empty product.

Examples 1.2.4:
( 1) x (1) = x

( 2) x (2) = x ( x + 1) = x2 + x

( 3) x (3) = x ( x + 1) ( x + 2 ) = x3 + 3 x2 + 2 x

( 4) x (4) = x ( x + 1) ( x + 2 ) ( x + 3) = x4 + 6 x3 + 11x2 + 6 x
Definition 1.2.5 [26]: The ‘Binomial Coefficients’ count the number
of ways to select k elements set out of n element set. The Binomial

n
Coefficients are denoted by   (Read as “n choose k”).
k 

 n
That is,
  = { T : T = { t , t ,..., t } ⊆ [ n] }
1 2 k

 k
Theorem 1.2.6(Binomial Theorem) [2]: Let n be a positive integer.
Then for all x and y we have,
n
n
( x + y ) n = ∑  xn y n −k
k =0  k 

Remark 1.2.7[2]: Binomial Coefficients also defined by,


n n!
  = r !( n − r ) ! , here n = 1, 2, …, n and r = 0,1,2,…,n.
k 

Definition 1.2.8[26]: The ‘Multinomial Coefficients’ count the


number of ways to split n distinct elements into k disjoint subsets, of

sizes n1 , n2 , …, nk respectively (of course, one assumes that


n1 + n2 +…+ nk = n ). The multinomial coefficients are denoted by 
 n1 , n 2

n 
.
,..., n k 
 n 
i.e.,   = { π : π = { T 1, T 2,..., Tk} , Ti ⊆ [ n ] , Ti = ni ,1 ≤ i ≤ k , n1 + n2 + ...nk = n} .
 n1 , n 2 ,..., n k 

Remark 1.2.9 [17]: Multinomial coefficients are also defined by


 n  n!
 =
 n1 , n 2 ,..., n k  n1!n2 !...nk !

Theorem1.2.10 (Multinomial Theorem) [17]: Let n be a positive

integer. Then for all x1 , x2 , …, xk we have,

 n  n1 n2 nk
( x1 + x2 + .... + xk ) ∑
n
=  x x ...xk
 n1 , n 2 ,..., n k  1 2
n1 + n 2 + ...nk = n

Theorem 1.2.11(The Principle of inclusion and exclusion) [6]: If


Ai are finite sub sets of universal set U, then the number of elements in


i =1
Ai is given by

n
A1 U A2 U ... U An = ∑ Ai − ∑ Ai I Aj + ∑ Ai I Aj I Ak + ... + ( −1)
n −1
A1 I A2 I ... I An ,
i =1 i, j i , j ,k

where the second summation is taken over all 2-combinations { i, j} of

the integers {1, 2, …, n}, the third summation is taken over a 3-

combinations { i, j , k } of {1, 2, …, n}, and so on.

Definition 1.2.12 (Gaussian binomial coefficient or q-binomial

n
coefficient) []: The ‘Gaussian binomial coefficient’ is denoted by  
 k q
and is defined as,
n (q n − 1)(q n −1 − 1)...(qn −k +1 − 1)
  = , for 0 ≤ k ≤ n and
 k q (q k − 1)(q k −1 − 1)...(q − 1)

n
  = 0 , otherwise.
 k q

Examples 1.2.13:
0
(1)   =1
 0 q

 1 q −1
(2)   = =1
 1 q q − 1

 2 q2 −1
(3)   = = 1+ q
 1 q q − 1

 3 q3 − 1
(4)   = = 1+ q + q2
1 q q − 1

3 (q 3 − 1)(q 2 − 1)
(5)   = 2 = 1 + q + q2
 2 q (q − 1)(q − 1)

 4 (q 4 − 1)(q3 − 1)
(6)   = = (q 2 + 1)(1 + q + q 2 ) = 1 + q + 2q2 + q3 + q4
 2 q (q − 1)(q − 1)
2

Properties1.2.13 []:

n  n 
(1)   =  
 k  q  n − k q

n k  n −1   n − 1
(2)   =q   + 
 k q  k q  k − 1q

 n   n −1  n −k  n − 1 
(3)   =  +q   , together with the boundary
 k  q  k q  k − 1q
conditions
n n
  =   = 1.
 0  q  n q

n  n  qn − 1
(4) =
    = = 1 + q + q 2 + ...q n −1
1
 q  n − 1 q q − 1

Definition 1.2.14[]: The ‘q-analogue’ of the natural number n by

[ n] q = 1 + q + q 2 + ...q n−1

Definition 1.2.15[]: The ‘q-factorial’ of the natural number is defined


by,

[ 0] q ! = 1 ,

[ n] q ! = [ 1] q [ 2] q ...[ n] q , for n ≥ 1.

Property1.2.16 []:

n [ n] q ! [ n] ![ n − 1] q !...[ n − k + 1] q !
  = = q
 k q [ k ] q ![ n − k ] q ! [ k ]q !

Theorem1.2.17 []: If q is a prime power, then the number of k-


dimensional subspaces of an n-dimensional finite vector space over a

n
finite field with q elements is   .
 k q

Theorem1.2.18 []: If q is a prime power, then the number of k −


dimensional subspaces of an n − dimensional finite vector space over
the finite field with q elements containing a given l − dimensional

n −l 
subspace is given by   .
 k − l q
Definition 1.2.19[26]: The ‘Stirling numbers of the second kind’
count the number of ways to partition a set of n elements into k
nonempty subsets (or blocks). Stirling numbers of the second kind are

n 
denoted by S ( n, k ) [5] or   [21] (read as “n subset k” or “n bracket
k 
k”),
n 
i.e.,   = { π : π = { T1 , T2 ,..., Tk } , Ti ⊆ [ n ] ,1 ≤ i ≤ k } , for 0 ≤ k ≤ n .
k 

4
Example 1.2.20:   = 7 , since there are seven different ways to
2

partition the set S = { 1, 2,3, 4} in to two non empty subsets. There are

(1). { { 1, 2,3} ,{ 4} }
(2). { { 1, 2, 4} , { 3} }
(3). { { 1,3, 4} ,{ 2} }
(4). { { 1,3, 4} ,{ 2} }
(5). { { 1, 2} , { 3, 4} }
(6). { { 1,3} ,{ 2, 4} }
(7). { { 1, 4} , { 2,3} }

Remark 1.2.21: The numbers n, k in definition 1.2.19 are assumed


non-negative integers. In spite of that, it is useful to simplify notation

n 
in all the necessary passages, giving a value to the number   even
k 

n 
when k < 0 . In this case, we conventionally put   = 0 . The following
k 
is a table for the first few values of stirling numbers of the second kind.
n\
k 0 1 2 3 4 5 6 7 8 9 10 11 12

0 1 0 0 0 0 0 0 0 0 0 0 0 0

1 0 1 0 0 0 0 0 0 0 0 0 0 0

2 0 1 1 0 0 0 0 0 0 0 0 0 0

3 0 1 3 1 0 0 0 0 0 0 0 0 0

4 0 1 7 6 1 0 0 0 0 0 0 0 0

5 0 1 15 25 10 1 0 0 0 0 0 0 0

6 0 1 31 90 65 15 1 0 0 0 0 0 0

7 0 1 63 301 350 140 21 1 0 0 0 0

8 0 1 127 966 1701 1050 266 28 1 0 0 0 0

9 0 1 255 3025 7770 6951 2646 462 36 1 0 0 0

10 0 1 511 9330 34105 42525 22827 5880 750 45 1 0 0


2850 14575
11 0 1 1023 1 0 246730 179487 63987 11880 1155 55 1 0
8652 61150 137940 132365 62739 15902 2227 170
12 0 1 2047 6 1 0 2 6 7 5 5 66 1
n 
Some initial values of  
k 

Properties1.2.22 []: Stirling numbers of second kind have the


following properties:
n 
(1)   = 0 if n < k , since there are no ways to partition a set of n
k 
elements into k non empty subsets if n < k .
n 
(2)   = 1 , if n = 0, k = 0 .
k 
n 
(3)   = 0 , if n > 0, k = 0 .
k 
n 
(4)   = 0 , if n = 0, k > 0 .
k 

Property1.2.23 [21]: The general formula for computing the Stirling


numbers of the second kind is
n  1 k i k
  = ∑ (−1)   ( k − i ) , for all positive integers n and k.
n

k  k ! i =0 i 
Proof: Finding an ordered partition of n elements into k blocks is

equivalent to finding onto functions from [ n ] to [ k ] . There are k n ways

to assign each of the distinct n elements a block. Then we subtract

k 
and add   ( k − i ) respectively until we have counted all possible onto
n

i 
functions, which is a standard application of the principle of inclusion-
exclusion and is exactly what the property states. This gives an

n 
explicit formula for   .
k 

Property1.2.24 (Recurrence relation) [21]:

Stirling numbers of the second kind obey the recurrence relation


given by
 n  n − 1  n −1
 = +  k , if n > 0 , k > 0 .
 k  k − 1  k 
Proof: Observe that a partition of the n objects into k nonempty

subsets either contains the subset { n} or it does not. The number of

ways that { n} is one of the subsets is given by


 n − 1
 .
 k − 1
Since we must partition the remaining n − 1 objects into the

available k − 1 subsets. The number of ways that { n} is not one of the

subsets (that is, n belongs to a subset containing other elements) is


given by
n −1
 k .
 k 

Since we partition all elements other than n into k subsets, and


then are left with k choices for inserting the element n.

Summing these two values gives the desired result.

Property1.2.25 [21]: The generating function for the Stirling


numbers of the second kind is

n
n 
∑ k  ( x) k = xn ,
k =0  

where ( x) k = ( x − 1)( x − 2)...( x − k + 1) is falling factorial polynomial.

Traditionally, this is the way in which Stirling numbers of second kind


are introduced.

n 
Property1.2.26 []:   = 1 , since there is only one way to partition a
1 
set of n elements into one block.

 n  n
Property1.2.27 []:   =   , this is because dividing n elements
 n −1  2 
into n − 1 sets necessarily means dividing it into one set of size 2 and
n−2 sets of size 1. Therefore we need only to pick those two
elements from n elements.
n 
Property1.2.27 []:   = 1 , since there is only one way to partition a
n 
set of n elements into n blocks.

Definition 1.2.38 (Bell numbers) [26]:The ‘Bell numbers’ are


n 
denoted by B (n) or Bn and defined as B (n) = ∑   , which counts all
k k 

n 
partitions of [ n ] , here   is a Stirling number of second .
k 

Examples1.2.35 []:
(1). B (3) = 5 , since there are five ways the numbers { 1, 2,3} can be
partitioned: { { 1} , { 2} , { 3} } , { { 1, 2} , { 3} } , { { 1,3} , { 2} } , { { 1} , { 2,3} } , and { { 1, 2,3} } .

(2). B (4) = 15 , since there are 15 ways the numbers { 1, 2,3, 4} can be

partitioned: { { 1} , { 2} ,{ 3} , { 4} } ,{ { 1, 2,3} , { 4} } ,{ { 1, 2, 4} ,{ 3} } ,
{ { 1,3, 4} { 2} } , { { 2,3, 4} { 1} } { { 1,3} ,{ 2, 4} } , { { 1, 4} , { 2,3} } , { { 1, 2} , { 3, 4} } , { { 1} , { 2} ,{ 3, 4} } ,
{ { 1} , { 3} , { 2, 4} } , { { 1} , { 4} ,{ 2,3} } , { { 2} ,{ 3} , { 1, 4} } , { { 2} , { 4} , { 1,3} } , { { 3} , { 4} ,{ 1, 2} } ,
and { { 1, 2,3, 4} } .
The diagram below shows the constructions giving B (3) = 5 and
B (4) = 15 , with line segments representing elements in the same subset
and dots representing subsets containing a single element.

B (3) = 5

B (4) = 15
Definition 1.2.36[]: The ‘sign less Stirling numbers of the first kind’
count the numbers of permutations of n elements with k disjoint
cycles. Singles Stirling numbers of first kind are denoted by c( n, k ) [] or

n 
 k  [] or S1 (n, k ) [].
 
n 
That is   = { σ : σ is the permutation of [ n ] with k disjoint cycles} , for k , n ∈ ¥ and
k 
1≤ k ≤ n .

Example 1.2.37[]:

(1) The list {1, 2, 3, 4} can be permuted into two cycles in the
following ways:
{{1,3,2},{4}}
{{1,2,3},{4}}
{{1,4,2},{3}}
{{1,2,4},{3}}
{{1,2},{3,4}}
{{1,4,3},{2}}
{{1,3,4},{2}}
{{1,3},{2,4}}
{{1,4},{2,3}}
{{1},{2,4,3}}
{{1},{2,3,4}}
4
There are 11 such permutations, thus   = 11 .
2

(2) Here are some illegible diagrams showing the cycles for
permutations of a list with five elements.

5
1  = 24 :
 

5 
3  = 35 :
 

5 
 4  = 10 :
 
5 
5  = 1 :
 

Remark 1.2.38: The following is a table for the first few values of sign
less Stirling numbers of the first kind.

n\k 0 1 2 3 4 5 6 7 8 9
0 1
1 0 1
2 0 1 1
3 0 2 3 1
4 0 6 11 6 1
5 0 24 50 35 10 1
6 0 120 274 225 85 15 1
7 0 720 1764 1624 735 175 21 1
8 0 5040 13068 13132 6769 1960 322 28 1
10958 11812 6728 2244
9 0 40320 4 4 4 9 4536 546 36 1

Definition 1.2.39[]: The ‘normal (or signed) Stirling numbers of the


5 is denoted by s (n, k ) [] or S ( k ) [] or S k [], and is denoted as,
first kind’
5 = 1 n n
 
n 
s (n, k ) = (−1) n − k   .
k 

4− 2  4 
Example 1.2.40: s (4, 2) = (−1)   = 11 .
2
Remark 1.2.41: The following is a table for the first few values of
normal Stirling numbers,

n\k 0 1 2 3 4 5 6 7 8 9
0 1
1 0 1
2 0 −1 1
3 0 2 −3 1
4 0 −6 11 −6 1
5 0 24 − 50 35 − 10 1
6 0 − 120 274 − 225 85 − 15 1
7 0 720 − 1764 1624 − 735 175 − 21 1
− 504 − 1313 − 196
8 0 0 13068 2 6769 0 322 − 28 1
− 10958 11812 − 6728 2244 − 453 −3
9 0 40320 4 4 4 9 6 546 6 1

Properties1.2.42 []: The following are some basic results on signed


and sign less Stirling numbers of first kind.

n
(1).   = 0 for n ≥ 1 .
0 

Since, there are no ways to arrange n objects in zero cycles.

n
(2).   = 1 for n ≥ 0 .
n

Since, there is only one way to assign n objects into n cycles, each
object in its own cycle.

n
(3).   = (n − 1)! , for n ≥ 1 .
1 

Since, if S = {σ : σ is a permutation on [ n ] with 1 cycle}


= {σ : σ is a cycle on [ n ] and its length n},

Then S = the number of n length cycles on [ n ]

= (n − 1)! .
Therefore by the definition 1.2.34, we have
n
1  = S
 
n
Hence,   = (n − 1)! .
1 

 n  n
(4).   = .
 n −1  2 

Since, if S = {σ : σ is a permutation on [ n ] with (n − 1) disjoint

cycles}, then it follows that every element is given its own cycle
except for one pair of elements, which is placed in a 2-cycle. Order is
irrelevant for entire case, so we only need to select the two elements

n
of the same cycle. Therefore we select   ways to select permutation
2
σ of S.
n
That is S =   .
2
By the definition 1.2.34, we have
 n 
 n −1 = S
 
 n  n
Hence,   = .
 n −1  2 

n 
(5).   = 0 if k > n .
k 
Since there are no ways to arrange n objects in k cycles if k > n .
(6). The generating function for stirling numbers of second kind is
given by
n
n 
∑  k x k
= ( x) n ,
k =0  
where, ( x) n = x ( x − 1)( x − 2)...( x − n + 1) is falling factorial polynomial.
That is, the signed stirling numbers of the first kind are the

coefficient of x k in the expansion of the falling factorial polynomial ( x) n


.
(7). The recurrence relation for stirling numbers of the first kind is
given by
 n   n − 1  n −1
 k  =  k − 1 + ( n − 1)  k  for all n ≥ k ≥ 1 ,
     
0  0 if k ≠ 0
(8).   = δ 0 k , where δ 0 k =  is kronecker delta.
k  1 if k = 0

Definition 1.2.42[]: A relation ≤ on a set L is said to be a ‘partial


order’ on L if it is reflexive, antisymmetric and transitive.
That is, (i) a ≤ a , for all a ∈ L (reflexive),
(ii) if a ≤ b and b ≤ a , then a = b , for a, b ∈ L
(antisymmetric), and
(iii) if a ≤ b and b ≤ c then a ≤ c , for a, b, c ∈ L (transitive).

Definition 1.2.43[]: If the relation ≤ on a set L is a partial order


then ( L, ≤) is called a ‘partially ordered set’ or simply a ‘poset’.

Definition 1.2.44[]: A partial order ≤ on a set L is said to be ‘total


order’ if for every pair of elements x, y ∈ L , we have either x ≤ y or y ≤ x
.
Definition 1.2.45[]: If the relation ≤ is a total on a set L , then we
say that ( L, ≤) is a ‘totally ordered set’ or ‘chain’.

Definition 1.2.46[]: A partially ordered set ( L, ≤) is called a ‘lattice’ if


every pair of elements in L has a least upper bound x ∨ y (x join y) and
greatest lower bound x ∧ y (x meat y).
That is , (i) For every x, y ∈ L , there is a unique element denoted by
x ∨ y called as least upper bound, with the property that z ≥ x and z ≥ y

u implies z ≥ x ∨ y , and
(ii) For every x, y ∈ L , there is a unique element denoted by
x ∧ y called as greatest lower bound, with the property that z ≤ x and
z ≤ y implies z ≤ x ∧ y .

Example1.2.47: L(G ) , the ‘lattice of subgroups’ of a group G is the


lattice whose elements are the subgroups of G, with the partial order
relation being set inclusion. In this lattice, the join of two subgroups is
the subgroup generated by their union, and the meet of two subgroups
is their intersection.

Definition 1.2.48[]: The ‘Hasse diagram’ (or ‘Lattice diagram’) of


partially ordered set (or lattice) L is the directed graph with vertex set
L and an edge form x up to y if y covers x (i.e., x < y and there is no z
such that x < z < y ).

Example1.2.49: The following diagrams are lattice diagrams of two


lattices namely, the devisors of 35 and the set of all subsets of three
element set { 1, 2,3} .
Definition 1.2.49[]: Let ( L, ≤) be a lattice, suppose there exists two
elements in y and z in L such that y ≤ x , for all x ∈ L and x ≤ z , for all
x ∈ L , then we say that y is the least element and z is the greatest
∧ ∧
element. The least and greatest elements usually denoted by 0 and 1

respectively.
CHAPTER II
MOBIUS INVERSION THEOREM OVER A
LATTICE OF SUBGROUPS

2.1. Definitions and some properties of action of


groups:
In this section, the definitions of group actions, stabilizer, orbit
and apereodic elements are given. Some basic properties are studies
from [].

Definition 2.1.1: Let G be a finite group with identity element e, and


S is a finite set. Then G is said to be acts on S if there is a mapping
φ : G × S → S , with φ( g , s ) = g ∗ s such that,

(i) g ∗ ( h ∗ s ) = ( gh) ∗ s , for all g , h ∈ G, s ∈ S and

(ii) e ∗ s = s , for all s ∈ S .

Example 2.1.2: Let ( G,.) be a finite group. Define g ∗ s = g .s ,

g ∈ G, s ∈ G . Then G acts on G it self.

Because, (i) g ∗ (h ∗ s ) = g *(h.s ) = g.( h.s ) = ( g.h).s = ( g.h) ∗ s , for all


g , h, s ∈ G ,

(ii) e * s = e.s = s , for all s ∈ G .

This action of the group G on itself is called translation.

Example 2.1.3: Let ( G,.) be a finite group. Define g ∗ s = g .s.g −1 ,

g ∈ G, s ∈ G . Then G acts on G it self.


Because, (i)

g ∗ ( h ∗ s ) = g *(h.s.h −1 ) = g.( h.s.h−1 ).g −1 = ( g .h).s.( g.h)−1 = ( g .h) ∗ s , for all g , h, s ∈ G .

(ii) e * s = e.s.e −1 = s , for all s ∈ G .

This action of the group G on itself is called conjugation.

Definition 2.1.4: Let G be a finite group with identity element e and

S is a finite set on which G acts. Given s ∈ S , the set Gs = { g ∈ G : g * s = s}

is called the stabilizer of s.

Lemma 2.1.5: Gs , the stabilizer of s, is a subgroup of G.

Proof: Clearly, Gs ≠ ∅ . Since e ∈ Gs , and Gs is a subset of finite group G.

Let g , h ∈ Gs . Then by using condition (i) of the definition 2.1.1,


we have,

( gh)* s = g *(h * s)

= g * s, since h ∈ Gs

= g , since g ∈ Gs .

Hence gh ∈ Gs .

This proves that Gs is a subgroup of G.

Definition 2.1.6: Let G be a finite group with identity element e and

S is a finite set on which G acts. Given s ∈ S , the set Os = { t ∈ S : t = g * s,

for some g ∈ G} is called the orbit of s.

Remark 2.1.7: Os , the orbit of s ∈ S , is a subset of S.


Theorem 2.1.8: Let G be a finite group with identity element e, and
let S be a finite set on wish G acts. Then

S = UOs (disjoint),
s∈C

where C is any sub set of S containing exactly one element from each
orbit.

Proof: For any r , s ∈ S , we define a relation ‘ : ’ on S by r : s iff r = g ∗ s

, for some g ∈ G .

Now by using (i) and (ii) conditions of definition 2.1.1, we have,

(1) s = e * s , for all s ∈ S .

Which imply that s : s for all s ∈ S .

Therefore the relation ‘ : ’ is reflexive.

(2) If r = g ∗ s , for some g ∈ G , then

g −1 * r = g −1 *( g * s)

= ( g −1 g ) * s

= e*s

= s.

That is, if r = g ∗ s , then s = g −1 * r , for some g ∈ G .

Which imply that, if r : s then s : r .

Therefore the relation ‘ : ’ is symmetric.


(3) If r = g ∗ s and s = h ∗ t , then

r = g *(h ∗ t )

= ( gh)* t , for any g , h ∈ G .

That is, if r : s and s : t , then r : t .

Therefore the relation ‘ : ’ is transitive.

Hence the relation ‘ : ’ is an equivalence relation on S, and hence


this equivalence relation partition S into mutually disjoint equivalence

classes. The equivalence class of an element s ∈ S is the orbit Os , the


orbit of s ∈ S . Because, the equivalence class of s ∈ S ,

[ s ] = { t ∈ S : t : s} = { t ∈ S :t = g * s, for some g ∈ G} = Os .

Hence, the set of all orbits is a partition of S.

Therefore, S = UOs (disjoint), where C is any sub set of S


s∈C

containing exactly one element from each orbit.

Theorem 2.1.9: Let G be a finite group with identity element e, and S


be a finite set on which G acts.

G
Then Os = , where | . |, denoted cardinality.
Gs

G
Proof: For a given s ∈ S , define a mapping φ : Os → by
Gs

φ( g * s ) = gGs , for all g ∈ G .

First we prove that φ is well defined.


Suppose, for any g , h ∈ G , g * s = h * s . Then by using (i) and (ii)

conditions of definition 2.1.1, we have ( g −1h) * s = g −1 *(h * s )

= g −1 *( g * s )

= e*s

= s.

Hence, g −1h ∈ Gs .

This shows that, gGs = hGs .

Therefore φ is well defined.

Now we prove φ is injective.

Suppose for any g , h ∈ G , we have φ( g * s ) = φ(h * s) . Then we get

gGs = hGs or g −1h ∈ Gs . That is ( g −1h) * s = s . Then in view of definition

2.1.1 we have,

g * s = g * ( g −1h ) * s 

=  g ( ( g −1h) )  * s

= ( eh ) * s

= h*s .

This shows that φ is injective.

Since, for any left coset gGs , there exists an element such that

g * s ∈ Os such that φ( g * s ) = gGs .


This shows that φ is surjective.

Therefore φ is bijective.

G
Hence, Os =
Gs

G
=
Gs
.

Definition 2.1.10: Let G be a finite group with identity element e,


and S is a finite set on which G acts. An element s ∈ S is said to be an
‘aperiodic element’ if the stabilizer of s ∈ S contains only e. That

is Gs = { e} .

Theorem 2.1.11: An element s ∈ S is aperiodic if and only if Os = Gs .

Proof: Suppose s ∈ S is aperiodic element. Then by definition 2.1.10

we have, Gs = { e} , that is Gs = 1 . Then by theorem 2.1.9 we obtain

Os = Gs .

Conversely suppose that Os = Gs , then by theorem2.1.9, we

obtain Gs = 1 , that is Gs = { e} , therefore s ∈ S is aperiodic element.

Theorem 2.1.12: The number of aperiodic elements in S is divisible

by G .

Proof: Let G be a finite group with identity element e and S is a finite


set on which G acts.
We first prove “If an element s ∈ S is aperiodic then so is g * s , for

every g ∈ G ”.

Suppose s ∈ S is an aperiodic element. Then by definition 2.1.10

we have Gs = { e} . That is g * s = s only if g = e .

Now h *( g * s) = (h.g ) * s = e only if h = e , so g * s is an aperiodic


element.

Hence if s ∈ S is aperiodic element then we get that every

element in Os is also aperiodic.

Hence if B = { s1 , s2 ,..., sn } is the set of all aperiodic elements of S

which are not related to each other then we get,

P = UOs (disjoint)
s∈B

Hence in view of theorem 2.1.11 we get

P = ∑ Os
s∈B

=∑G
s∈B

=nG
.

Which prove that number of aperiodic elements in S is divisible by G .

Definition 2.1.13: Let L(G ) be the lattice of subgroups of G under

partial ordering by inclusion. For H ∈ L(G ) , α ( H ) is the number of


elements in S, which are stabilized by H.
That is, α ( H ) = { s ∈ S : Gs = H } .

Theorem 2.1.14: α ( { e} ) ≡ 0 ( mod G ) .

Proof: By using definition 2.1.13 we have,

α ({ e} ) = { s ∈ S : Gs = { e} }

i.e., α ( { e} ) denotes the number of aperiodic elements in S.

Therefore, by using theorem 2.1.12 we have, | G | α ( { e} )

i.e., α ( { e} ) ≡ 0 ( mod G ) .

Definition 2.1.15: Let L(G ) be the lattice of subgroups of G under

partial ordering by inclusion. For H ∈ L(G ) , β ( H ) be the number of

elements in S, whose stabilizer contains H.

That is, β ( H ) = { s ∈ S : Gs ⊇ H } .

Remark 2.1.16: It is obvious that β ( H ) ≥ α ( H ) .

Theorem 2.1.17 (The relation between α ( H ) and β ( H ) ):

β(H) = ∑
K ∈L (G )
α (K )
K ≥H

Proof: Let
Α= U { s∈S :G s = K}
and B = { s ∈ S : Gs ⊇ H }
K ∈L ( G )
K ≥H
We now show that A = B .

Suppose, t ∈ A , then t ∈ S and Gt = K , for some K ∈ L(G ) , K ≥ H .

i.e., t ∈ S and Gt ⊇ H

i.e., t ∈ B .

Conversely suppose that t ∈ B .

Then t ∈ S and Gt ⊇ H

i.e., t ∈ S and Gt = K , for some K ∈ L(G ) , K ≥ H .

i.e., t ∈ A .

Hence A=B
(1)

We now show that the union defined in A is disjoint.

Suppose Ai = { s ∈ S : Gs = Ki } and A j = { s ∈ S : Gs = K j } , for some

K i , K j ∈ L(G ) and K i ≥ H , K j ≥ H , with i ≠ j .

If possible assume that t ∈ Ai I Aj

i.e., t ∈ Ai and t ∈ A j

i.e., t ∈ S , Gt = Ki and Gt = K j , for some K i , K j ∈ L(G ) and K i ≥ H , K j ≥ H .

i.e., K i = K j , for some K i ,K j ∈L (G ) and K i ≥ H , K j ≥ H

Which is a contradiction, since i ≠ j .


Hence Ai I Aj = ∅ , for i ≠ j.

(2)

Therefore the union defined in A is disjoint.

By using the results (1) and (2) we obtain,

β ( H ) = { s ∈ S : Gs ⊇ H }

= B

= A

= U { s∈S :G
K ∈L (G )
s = K}
K ≥H

= ∑ { s∈S :G
K ∈L ( G )
s = K}
K ≥H

= ∑ α(H)
K ∈L ( G )
K ≥H

2.2. Möbius inversion theorem over a lattice of


subgroups:
In this section we define Möbius function on lattices and we
prove the Möbius inversion theorem over a lattice of subgroups which
∧ ∧
is given in []. Throughout this section 0 is the minimal element and 1

is maximal element and the relation H ≤ G means H is a subgroup of G.


Definition 2.2.1(Möbius function): Let L be a finite lattice with
∧ ∧
unique minimal 0 element, unique maximal element 1 and Möbius

function be the set of all integers. The Möbius function µ : L → ¢ is


defined recursively by,

(i) µ ( a ) = 1 , if a = 0 and

(ii) µ ( a ) = −∑ µ ( b ) , for

b<a a ≠ 0.

Theorem 2.2.2: The Möbius function µ : L → ¢ is the unique ¢ −

valued function defined on L such that ∑ µ ( b) = δ


b≤a

0a
, where δ 0a
∧ is

kronecker delta.

Proof: By definition 2.2.1, and by using the fact that 0 is the minimal

element of L, we get If a = 0 , then


∧ ∑ µ (b) = ∑ µ (b)
b≤a ∧
b≤0

= ∑ µ (b )

b =0


= µ (0)
=1.

If a ≠ 0 , then ∑ µ (b) = ∑ µ (b) + ∑ µ (b)
b≤a b =a b<a

= µ ( a) − µ (a )
=0

Hence, ∑ µ ( b) = δ
b≤a

0a
, where δ 0a
∧ is kronecker delta.

Theorem 2.2.3: Let µ : L → ¢ be a Möbius function of a finite lattice

( L, ≤) and let a ∈ L with a > 0∧ . Then,

∑ µ ( x ) = 0 , where

∧ ∧
0 and 1 are minimal and maximal
x:x ∨ a =1

elements of L respectively.
Proof: By using theorem 2.2.2, we have

∑ µ ( x ) = ∑ 1.µ ( x )
∧ x∈L
x:x ∨ a =1 ∧
x ∨ a =1

= ∑1. ∑ µ ( x )
x∈L ∧
x ∨ a =1

= ∑1.∑ µ ( x ) ∧ ∧

x∈L ∧ , since 0 ≠ 1
x ≤1

=0.
Theorem 2.2.4: Let µ : L → ¢ be a Möbius function of a finite lattice


( L, ≤) and let a, b ∈ L such that 0 < a ≤ b in L, then

µ ( b) = − ∑ µ ( c)
c <b
c ∨ a =b

Proof: By using definition 2.2.1, for b ≠ 0 we have,


µ ( b ) = −∑ µ ( c )
c <b

=− ∑ µ ( c)
c <b ∧
c ∨ a =b
, since 0 < a ≤ b .

Theorem 2.2.5(Möbius inversion theorem over a lattice): Let ( L, ≤) be a


lattice with unique minimal element 0 , a ∈ L and let ¢ be the set of all

integers. Suppose there are two functions α : L → ¢ and β : L → ¢


satisfying the condition

β (b) = ∑ α (a )
a ≥b , for all b ∈ L ,
(1)
α  0  = ∑ µ ( b ) β (b )

then,   b∈L ,
(2)

where µ is a Möbius function defined on L.

Proof: By using (1) equation of 2.2.4 and 2.2.2 we obtain,

 
∑ µ ( b ) β (b) = ∑ µ ( b ) ∑ α (a ) 

b∈L b∈L
 b≤ a 
 a ,b∈L 

= ∑ µ ( b )α (a)
b≤a
a ,b∈L

= ∑ α (a ) µ ( b )
b≤a
a ,b∈L

 
= ∑ α (a ) ∑ µ ( b ) 

a∈L
 b≤ a 
 a , b∈L 

   
= ∑ α (a ) ∑ µ ( b ) + ∑ α ( a) ∑ µ ( b ) 
  
a∈L
 b≤ a  a∈L  b≤ a 

a =0
 a ,b∈L  ∧
a ≠0
 a ,b∈L 

∧
= α  0  .1 + 0
 

∧
= α 0
 
Theorem 2.2.6(Möbius inversion theorem over a lattice of subgroups):

Let L(G ) be the lattice of subgroups of a group G, under the partial

order relation being set inclusion, with unique minimal element


{ e} ,
H ∈ L(G ) and let ¢ be the set of all integers. Suppose there are two

functions α : L(G ) → ¢ and β : L(G ) → ¢ satisfying the condition

β (H ) = ∑
K ≥H
α (K )
K ∈L ( G )
,

α ( { e} ) = ∑ µ ( H ) β (H )
then H ∈L (G )
,

where µ : L(G ) → ¢ is a Möbius function.

Proof: Since, L(G ) be the lattice of subgroups of a group G, under the


partial order relation being set inclusion, with unique minimal element

{ e} , so we apply theorem 2.2.5 to this lattice we get required theorem.

∑ µ ( H ) β (H ) ≡ 0 (mod G )
Corollary 2.2.7: H ∈L (G )
.

Proof: By theorem 2.1.14 we have,

α ( { e} ) ≡ 0 ( mod G )
.

And by the theorem 2.2.6 we have,

α ( { e} ) = ∑ µ ( H ) β (H )
H ∈L ( G )
.

∑ µ ( H ) β (H ) ≡ 0 (mod G )
Hence we get, H ∈L (G ) .
This proves the corollary.

CHAPTER III

SOME CONGRUENCES ON BINOMIAL


COEFFICIENTS AND STIRLING NUMBERS OF
BOTH KINDS IN THE CASE G ═ Cp
3.1 Möbius function defined on L(C P) :
In this section, we find Möbius function defined on L(CP ) , and hencewe prove a
congruence relation from corollary 2.2.7.
Lemma 3.1.1: The Möbius function µ defined on L(CP ) is given by,

 1 if H = { e}
µ (H ) = 
 −1 if H = CP

Proof: Let G = CP acts on S, it is obvious that L(CP ) = { { e} , CP } , and ( L(CP ), ⊆ ) is a


lattice and its lattice diagram is,

Since { e} is a unique minimal element in L(CP ) , then by the definition 2.2.1, we have

µ ( { e} ) = 1 .

Also since, CP is the only element of L(CP ) which is not equal to { e} we get in view of
the definition 2.2.1 that,
µ (CP ) = − ∑ µ ( H )
H < CP
= − µ ( { e} )

 1 if H = { e}
Hence we get , µ (H ) =  .
−1 if H = CP

Theorem 3.1.2: If G = CP acts on S, then

β ( { e} ) ≡ β ( CP ) (mod p ) .

Proof: In view of corollary 2.2.7 we get,


H ∈L (C P )
µ ( H ) β ( H ) ≡ 0 (mod CP )

i.e., µ ( { e} ) β ( { e} ) + µ (CP ) β ( CP ) ≡ 0 (mod p ) ,


Using lemma 3.1.1 we get
β ( { e} ) − β ( CP ) ≡ 0 (mod p )

i.e., β ( { e} ) ≡ β ( CP ) (mod p ) .

3.2. Congruence relation on Binomial coefficients in the case


G ═ Cp :
In this section, we prove congruence relation [] on binomial coefficients in the

case G = CP by using theorem 3.1.2.

Theorem 3.2.1: Let S be the set of all k element sub sets T of [ n + p ] and let CP acts on

S in a natural way, i.e., if g ∈ CP and T = { t1 , t2 ,..., tk } , then g * t = { g * t1 , g * t2 ,..., g * tk } .


Then
n+ p n  n 
  ≡ +  ( mod p )
 k  k  k − p

Proof: Let S = { T : T = { t1 , t2 ,..., tk } ⊆ [ n + p ] } . Then by the definition of binomial

coefficients,
n+ p
S =  (1)
 k 
Let G = CP acts on S in natural way. If every k-sub set T of [ n + p ] is stabilized by { e} ,

that is GT = { e} , ∀T ∈ S , then by theorem 2.1.17, definition 2.1.13and (1) we obtain,

β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP )

= ∑ {T ∈S :G
K ≥{ e}
T = K}
K ∈L ( CP )

= { T ∈ S : GT = { e} } + { T ∈ S : GT = CP }

= S + ∅ , since GT = { e} , ∀T ∈ S

= S

n+ p
=  (2)
 k 

Now if every k-sub set T of [ n + p ] is stabilized by CP , that is GT = CP , ∀T ∈ S ,


there are two possibilities.

If 1 ∈ T , then the repeated action of (1, 2,..., p) forces [ p] ⊆ T , that is

{ 1, 2,..., p} ⊆ T = { t1 , t2 ,..., tk } .
Now we must chose the k − p elements of T out of the elements of
p + [ n ] = { p + 1. p + 2,..., p + n} .

 n 
Thus we get   ways for T.
k − p

Hence, { T ∈ S : GT = CP ,1 ∈ T }

{
= T ∈ S : T = { 1, 2,..., p, t1 , t2 ,..., tk − p } and { t1 , t2 ,..., tk − p } ⊆ p + [ n ] }
 n 
=  (3)
k − p
If 1 ∉ T , then we have T ⊆ p + [ n ] , i.e., T = { t1 , t2 ,..., tk } ⊆ { p + 1. p + 2,..., p + n} .

Now we choose k elements of T out of n elements of p + [ n ] .

n
Thus we get   ways for T.
k 

Hence, { T ∈ S : GT = CP ,1 ∉ T }

= { T ∈ S : T = { t1 , t2 ,..., tk } ⊆ p + [ n ] }
n
=  (4)
k
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,

β ( CP ) = ∑
K ≥C P
α (K )
K ∈L ( CP )

= ∑ {T ∈S :G
K ≥ CP
T = K}
K ∈L ( CP )

∑ {T ∈S :G T = K} + ∑ {T ∈S :G T = K}
= K ≥CP K ≥ CP
K ∈L ( CP ) K ∈L (CP )
1∈T 1∉T

= { T ∈ S : GT = CP ,1 ∈ T } + { T ∈ S : GT = CP ,1 ∉ T }

 n  n
= +  (5)
k − p k 

Hence in view of theorem 3.1.2, equations (2) and (5) we get


n + p n  n 
  ≡ +  ( mod p ) .
 k  k  k − p
3.3. Congruence relations on Stirling numbers of second kind
in the case G ═ Cp:
In this section we prove Congruence relation [] on Stirling numbers of second

kind in the case G = C p by using theorem 3.1.2. Using this congruence relation we
shall prove some theorems from [] , [] , [] , and [] .

Theorem 3.3.1: Let S be the set of all partitions π of [ n + p ] into k subsets (or blocks),

n + p 
which are counted by the stirling numbers of the second kind   . The action of CP
 k 
on subsets naturally extends to such partitions. Then,
 n + p   n  n + 1
 ≡ +   (mod p )
 k  k − p   k 

Proof: Let S = { π : π = { T1 , T2 ,..., Tk } , Ti ⊆ [ n + p ] ,1 ≤ i ≤ k } and by the definition of

Stirling numbers of second kind, we have


n + p 
S = 
 k 
(1)
Let G = CP acts on S in natural way, that is if g ∈ CP and π = { T1 , T2 ,..., Tk } ∈ S ,

then g * π = { g * T1 , g * T2 ,..., g * Tk } .

If every k-partition π of [ n + p ] is stabilized by { e} , that is Gπ = { e} , ∀π ∈ S ,


then by theorem 2.1.17, definition 2.1.13and (1) we obtain,
β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP )

= ∑ { π∈ S :G
K ≥{ e}
π = K}
K ∈L ( CP )

= { π ∈ S : Gπ = { e} } + { π ∈ S : Gπ = CP }

= S + ∅ , since Gπ = { e} , ∀π ∈ S

= S

n + p 
=  (2)
 k 
Suppose for the partitions π if CP stabilizes, that is Gπ = CP , ∀π ∈ S , then to

compute β (CP ) there are two possibilities.

If { 1} ∈ π then { 1} is a singleton block of π , then so are { 2} , { 3} ,..., { p} . That is

{ { 1} , { 2} , { 3} ,...,{ p} } ⊆ π = { T , T ,..., T } ,
1 2 k where Ti ⊆ [ n + p ] ,1 ≤ i ≤ k . Now, we

partitioning the remaining k-p blocks of π from the set p + [ n ] . This means there are

 n 
  ways to chose the rest of the blocks of π .
k − p 

Thus, { π ∈ S : Gπ = CP , { 1} ∈ π}

{
= π ∈ S : π = { { 1} .{ 2} , { 3} ,..., { p} , T1 , T2 ,..., Tk − p } and { T1 , T2 ,..., Tk − p } ⊆ p + [ n] }
 n 
=  (3)
k − p 

If { 1} ∉ π , then 1 is in a block of size at least two, so 2,3,….,p are also in the same

block. Thus the set [ p ] = { 1, 2,..., p} is acting like a single element, which we shall

denote by P, so we are really just partitioning the set { P, p + 1, p + 2,..., p + n} into k-

 n + 1
blocks. This means there are   ways to choose the blocks of π .
 k 
Thus, { π ∈ S : Gπ = CP , { 1} ∉ π}

= { π ∈ S : π = { T1 , T2 ,..., Tk } and Ti ⊆ { P, p + 1, p + 2,..., p + n} ,1 ≤ i ≤ k }

 n + 1
=  (4)
 k 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,
β ( CP ) = ∑
K ≥C P
α (K )
K ∈L ( CP )

= ∑ { π∈ S :G
K ≥ CP
π = K}
K ∈L ( CP )

= ∑ { π∈ S :G
K ≥ CP
π = K} + ∑ { π∈ S :G
K ≥CP
π = K}
K ∈L ( CP ) K ∈L ( CP )
{ 1} ∈π {1}∉π

= { π ∈ S : Gπ = CP , { 1} ∈ π} + { π ∈ S : Gπ = CP , { 1} ∉ π}

 n  n + 1
= +   (5)
k − p   k 
Hence in view of theorem 3.1.2, equations (2) and (5) we get
n + p   n  n + 1
 ≡ +   (mod p )
 k  k − p   k 

Using this congruence relation we can prove the following theorems.

Theorem 3.3.2 (Lagrange []): ( x) p = x p − x (mod p ) , where ( x) p is the falling

factorial polynomial.
Proof: By 1.2.25, we have the generating function for the stirling numbers of the second
kind,

n
n 
∑ k  ( x) k = xn ,
k =0  

Where ( x) k = ( x − 1)( x − 2)...( x − k + 1) is falling factorial polynomial.


Taking n = p in 1.2.25 we get,

p
 p
∑ k  ( x) k = xp (1)
k =0  

Setting n = 0 in theorem 3.3.1, we get

 p   0  1 
 ≡  +   (mod p ) (2)
 k  k − p  k 

From (1) and (2) and by using 1.2.2 and 1.2.22 we get,

p
 p p
  0  1  
x p = ∑   ( x) k ≡ ∑    +   ( x)k (mod p)
k =0 k  k =0  k − p  k  

  0  1     0  1    0  1  
≡   +    ( x)0 +    +    ( x)1 + ... +    +    ( x) p (mod p)
 0 − p  0    1 − p  1   0   p  

≡ ( 0 + 0 ) ( x)0 + ( 0 + 1) ( x )1 + ... ( 1 + 0 ) ( x) p (mod p )

≡ ( x)1 + ( x ) p (mod p )

≡ x + ( x) p (mod p )

i.e., x − x ≡ ( x ) p (mod p) .
p

Theorem 3.3.3(Joe De Maio []):If p is prime then,


 p
  ≡ 0 (mod p ), for all 2 ≤ k ≤ p − 1 .
k 
 p  p
Furthermore, p Œ
 ,   .
1   p 

Proof: Taking n = 0 in the theorem 3.3.1, we get

 p   0  1 
 ≡  +   (mod p ) (1)
 k  k − p  k 
Since, 2 ≤ k ≤ p − 1 , then by using remark 1.2.21 and property (1) of 1.2.22 we get,
 0  1 
  = 0 and   = 0 .
k − p  k 
Using these values in (1) we get ,
 p
  ≡ 0 (mod p )
k
Moreover, taking k = 1 in (1), and by using remark1.2.21 and property 1.2.27 we get,
 p   0  1
 ≡  +   (mod p )
1  1 − p  1
≡ 0 + 1 (mod p)
≡ 1 (mod p ) .

 p
Thus,   ≡ 1 (mod p ) ,
1 
 p
i.e., p Œ
 .
1 
Now, taking k = p in (1) and by using (2) and (1) properties of 1.2.22 we get
 p  0  1 
  ≡   +   (mod p )
 p  0   p 
≡ 1 + 0 (mod p )
≡ 1 (mod p ) .

p
Thus,   ≡ 1 (mod p ) ,
 p
 p
i.e., p Œ
 .
 p
Theorem 3.3.4(Joe De Maio []):If p is a prime then,
 p + 1
  ≡ 0 (mod p), for all 2 ≤ k ≤ p − 1 .
k + 1 
 p +1
Furthermore,   ≡ 1 (mod p ).
 2 
Proof: Replacing k by k+1 and n by 1 in the theorem 3.3.1, we get
 p + 1  1  1 + 1 
 ≡ +   (mod p)
 k + 1  (k + 1) − p   k + 1
 1   2 
≡ +   (mod p ) (1)
(k − p ) + 1 k + 1
Since, 2 ≤ k ≤ p − 1 , then by using remark 1.2.21 and (1) and (3) property of 1.2.22 we

 2   1 
get,   = 0 and   = 0.
 k + 1 (k − p) + 1
Using these values in (1) we get ,
 p + 1
  ≡ 0 (mod p) .
k + 1 
Moreover, taking k = 1 in (1) and by using property (3) of 1.2.22, remark 1.2.21 and
property 1.2.27 we get,
 p + 1  1  2 
 ≡  +   (mod p )
 2   2 − p  2 
≡ 0 + 1 (mod p)
≡ 1 (mod p )

 p + 1
Thus,   ≡ 1 (mod p ) .
 2 
Theorem 3.3.5 (Joe De Maio []):If p is a prime then,
 p + j
  ≡ 0 (mod p ), for all 1 ≤ j ≤ p − 2 and 2 ≤ k ≤ p − j .
k + j 

 p + j
Further more,   ≡ 1 (mod p ), for all 2 ≤ j ≤ p − 2 .
 j +1 
Proof: Replacing k by k + j and n by j in the theorem 3.3.1, for all 1 ≤ j ≤ p − 2 and
2 ≤ k ≤ p − j we get,

 j + p  j +1  j 
 ≡ +   (mod p ) (1)
 k + j  k + j   k + j − p 
Since, 2 ≤ k ≤ p − j , i.e., j + 2 ≤ k + j ≤ p ,
Therefore, j + 1 < k + j , and hence by using property (1) of 1.2.22, weget
 j +1
 =0 (2)
k + j 
Since, 2 ≤ k ≤ p − j , i.e., k + j − p ≤ 0 , and since, 1 ≤ j ≤ p − 2 , i.e., j ≠ 0 , then by
using Remark 1.2.21 we get,
 j 
 =0 (3)
 k + j − p
Using the values (2) and (3) in (1) we get,
 p + j
  ≡ 0 (mod p ), for all 1 ≤ j ≤ p − 2 and 2 ≤ k ≤ p − j .
k + j 
Moreover, taking k = 1 in (1) and by using property 1,2,27 and remark 1.2.21 we get,
 p + j   j + 1  j 
 ≡ +   (mod p )
 j + 1   j + 1  j + 1 − p
≡ 1 + 0 (mod p ) , since, j ≠ 0 and j + 1 < p
≡ 1 (mod p ) .
Theorem 3.3.6 (Joe De Maio []): If n is a composite number then there exists k,

n 
2 ≤ k ≤ n − 1 such that n Œ
 .
k 
Proof: Suppose n is a composite number. Let p be any prime factor of n and let
j = n − p . Since n is composite, then 2 ≤ n − p ≤ n − 2 , that is 2 ≤ j ≤ n − 2 .
By theorem 3.3.5, we have
 p + j
  ≡ 1 (mod p ) , for all 2 ≤ j ≤ n − 2 (1)
 j +1 
Thus, taking j = n − p in (1), we get

 p + ( n − p ) 
  ≡ 1 (mod p ), since, 2 ≤ n − p ≤ n − 2
( n − p ) + 1 

 n 
i.e.,   ≡ 1 (mod p ),
( n − p ) + 1
n 
i.e.,   ≡ 1 (mod p ), for k = ( n − p ) + 1 , 2 ≤ k ≤ n − 1 .
k 
n 
  , for 2 ≤ k ≤ n − 1 .
Hence, n Œ
k 
Due to theorem 3.3.3 and theorem 3.3.6 can now be improved the following
theorem.
Theorem 3.3.7 (Joe De Maio []): The positive integer n is a prime number iff

n 
  ≡ 0 (mod p ) for 2 ≤ k ≤ n − 1 .
k 
Proof: Suppose n is a prime number, then by theorem 3.3.3 we have
n 
  ≡ 0 (mod p ) , for all 2 ≤ k ≤ p − 1 .
k 
n 
Conversely suppose that,   ≡ 0 (mod p ) , for all 2 ≤ k ≤ p − 1 .
k 
If possible assume that n is a composite number then by theorem 3.3.6 we have,
n 
  , for 2 ≤ k ≤ n − 1

k 
n 
i.e.,   ≡/ 0 (mod p ) .
k 
This is a contradiction to our assumption.
Hence, n must be a prime number.
Theorem 3.3.6 (Touchard []):
B (n + p) ≡ B (n + 1) + B( n) (mod p )
Proof: By theorem 3.3.1, for each k we have
 n + p   n  n + 1
 ≡ +   (mod p ) ,
 k  k − p   k 
By summing these congruence for all values of k we get,
n + p   n + 1  n 
∑  ≡∑  +∑  (mod p)
k  k  k  k  k k − p 
By using definition 1.2.34 we get,
B (n + p) ≡ B (n + 1) + B( n) (mod p ) .

Becker and Riordan [] proved the following theorem by induction on j through the
repeated application of theorem 3.3.1. There is also a way to derive this congruence
relation directly by using group actions.

Theorem 3.3.7(Becker and Riordan[]): If j is defined by the equation p j ≤ k ≤ p j +1


then,
 n + p j ( p − 1)   n 
  ≡   (mod p )
 k  k 

{ }
Proof: Let S = π : π = { T1 , T2 ,..., Tk } , Ti ⊆  n + p  ,1 ≤ i ≤ k , where j is defined by the
j +1

equation p j ≤ k ≤ p j +1 .
By using the definition 1.2.19 we have,
n + p j +1 
S = 
 k 
(1)

Let G = C p j +1 acts on S in natural way.

If every k-partition π of  n + p  is stabilized by { e} , that is Gπ = { e} , ∀π ∈ S , then by


j +1

theorem 2.1.17, definition 2.1.13and (1) we obtain,

β ( { e} ) = ∑ α (K )
K ≥{ e}
(
K ∈L C
p j +1 )
= ∑ { π∈ S : G
K ≥{ e}
π = K}

(
K ∈L C
p j +1 )
= { π ∈ S : Gπ = { e} } + π ∈ S : Gπ = C{ p j+ 1 }
= S + ∅ , since Gπ = { e} , ∀π ∈ S
= S

n + p j +1 
=  (2)
 k 

Suppose for the partitions π if C p j +1 stabilizes, that is if Gπ = C p j +1 , ∀π ∈ S , then

to compute β (CP ) there are two possibilities.

If the set  p  acts a single element, which we shall denote by P, then we are
j +1

partitioning the set { P, p + 1, p + 2,..., p + n} into k- blocks. This means there are
j +1 j +1 j +1

 n + 1
  ways to choose the k blocks of π .
 k 

{
Thus, π ∈ S : π = { T1 , T2 ,..., Tk } and Ti ⊆ { P, p + 1, p + 2,..., p + n} ,1 ≤ i ≤ k
j +1 j +1 j +1
}
n + 1
=  (3)
 k 

If the set  p  is partitioned into p i sets of size p j +1−i for 1 ≤ i ≤ j + 1 , i.e.,


j +1

 n 
Tr' = { r + sp i : 0 ≤ s ≤ p j +1 −i − 1} for 1 ≤ i ≤ j + 1 , then there are  i  ways to choose
k − p 
the blocks of π for 1 ≤ i ≤ j + 1 .

Thus, { π ∈ S : π = { T , T ,..., T , T , T
1
'
2
' '
pi pi p i +1 }
,..., Tk and Ti ⊆ p j +1 + [ n ] , pi ≤ i ≤ k }
 n 
= i  ,for
1 ≤ i ≤ j +1 (4)
k − p 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,

(
β C p j+1 = ) ∑
K ≥C
α (K )
p j +1

(
K ∈L C
p j +1 )
∑ { π∈ S : G π = K}
= K ≥C
p j +1

(
K ∈L C
p j +1 )
n + 1 j +1  n 
= + ∑ i (5)
 k  i =1  k − p 
From theorem 2.1.8 and 2.1.9 we see that every element of S lies in an orbit
whose size is either divisible by p or equal to 1, and hence

(
β ( { e} ) ≡ β C p j +1 ) (mod p ) (6)

From (2), (5) and (6) we have,


 n + p j +1  n + 1 j +1  n 
 ≡  + ∑ i
(mod p)
 k   k  i =1  k − p 
 n 
But if k < p j +1 then  j +1 
= 0 , and hence we get
k − p 
 n + p j +1  n + 1 j  n 
 ≡  + ∑ i
(mod p)
 k   k  i =1  k − p 
n + p j 
≡  (mod p)
 k 
 n + p j +1 − p j  n 
i.e.,   ≡   (mod p )
 k  k
 n + p j ( p − 1)   n 
i.e.,   ≡   (mod p )
 k  k

3.4. Congruence relations on Stirling numbers of first kind in


the case G ═ Cp:
In this section we prove congruence relation [] on Stirling numbers of second kind

in the case G = C p by using theorem 3.1.2. Using this congruence relation we shall
prove Wilson’s theorem [] and some other congruence relation from [] .
Theorem 3.4.1: Let S be the set of all permutations σ of [ n + p ] into k disjoint cycles,

n + p 
which are counted by the stirling numbers of the first kind   . Let CP acts on S by
 k 
conjugation. Then,
n + p   n   n 
 k  ≡  k − p  +( p − 1)  k − 1 (mod p ) .
     

Proof: Let S = { σ : σ is a permutation of [ n + p ] with k disjoint cycles} and by the

definition of stirling numbers of first kind, we have


n + p 
S =  (1)
 k 
Let G = CP acts on S by conjugation, that is if g ∈ CP and σ ∈ S (i.e., σ = τ 1τ 2 ...τ k ,

where τ i is a cycle of length ni , 1 ≤ i ≤ k , with n1 ≤ n2 ≤ ... ≤ nk and n1 + n2 + ... + nk

= n + p ) then g * σ = gσg −1 = g(τ 1τ 2 ...τ k )g−1 = (gτ1g−1 )(gτ 2 g−1 )...(gτ k g−1 ) .

If every permutation σ of [ n + p ] is stabilized by { e} , that is Gσ = { e} , ∀σ ∈ S ,


then by theorem 2.1.17, definition 2.1.13and (1) we obtain,

β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP )

= ∑ { π∈ S :G
K ≥{ e}
π = K}
K ∈L ( CP )

= { σ ∈ S : Gσ = { e} } + { σ ∈ S : Gσ = CP }

= S + ∅ , since Gσ = { e} , ∀σ ∈ S

= S

n + p
=  (2)
 k 
Suppose Gσ = CP , ∀σ ∈ S , then to compute β (CP ) there are two possibilities.
If 1 is in a cycle τ of length 1, then ( 1) is the one cycle of σ and so

( 2 ) , ( 3) ,..., ( p ) are also cycles of the product of cycles σ . Now, we must choose the

 n 
remaining k − p cycles of σ from the set p + [ n ] . This means there are   ways to
k − p 
chose the rest of the cycles of σ .

Thus, { σ ∈ S : Gσ = CP ,1 ∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S }

= { σ ∈ S : σ = ( 1) ( 2 ) ... ( p ) τ 1τ 2 ...τ k − p and τ i is a cycle on p + [ n ] ,1 ≤ i ≤ k − p }

 n 
=   (3)
k − p 

If 1 is in a cycle τ of length at least two, then 2,3,….,p are also in the same

cycle, i.e., τ = (1, 2,3,..., p) . We chose remaining k − 1 cycles of σ from the set p + [ n ] .

 n 
This means there are   ways to construct the rest of the cycles of σ , and also there
 k − 1

are p − 1 choices for τ , i.e., τ = (1, 2,3,..., p) j , for 1 ≤ j < p . Therefore there are

 n 
( p − 1)   choices for cycles of σ .
 k − 1

Thus, { σ ∈ S : Gσ = CP ,1 ∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }

= { σ ∈ S : σ = ττ 1τ 2 ...τ k − p and τ i is a cycle on p + [ n ] ,1 ≤ i ≤ k − p }

 n 
= ( p − 1)   (4)
 k − 1
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,
β ( CP ) = ∑
K ≥C P
α (K )
K ∈L ( CP )

= ∑ { σ∈ S :G
K ≥ CP
σ = K}
K ∈L ( CP )
= { σ ∈ S : Gσ = CP ,1 ∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S } +

{ σ ∈ S : Gσ = CP ,1∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }

 n   n 
=  +( p − 1)   (5)
k − p   k − 1
Hence in view of theorem 3.1.2, equations (2) and (5) we get,
n + p   n   n 
 k  ≡  k − p  +( p − 1)  k − 1 (mod p )
     

Using this congruence relation we can prove the following theorems.

Theorem 3.4.2(Wilson’s theorem):


If p is a prime number then, ( p − 1)! ≡ −1 ( mod p ) .
Proof: Put n = 0 and k = 1 in theorem 3.4.1 we get,
 p  0   0
1  ≡ 1 − p  +( p − 1)  0  (mod p ) (1)
     
By (3) of property 1.2.34, we get
 p
1  = ( p − 1)! (2)
 
And by (8) of property 1.2.34, weget
 0  0 if k ≠ 0
 k  = 1 if k = 0
  
0   0 
Therefore we get,   = 1 and   = 0 (since, p ≠ 1 ).
0  1 − p 
Using these values and equation (2) in (1) we get,
( p − 1)! ≡ p − 1 ( mod p )

i.e., ( p − 1)! ≡ −1 ( mod p ) , since p − 1 ≡ −1 ( mod p ) .


 n
Theorem 3.4.3[]:   ≡ 0 (mod p ) if n > kp , where p is a prime number.
k 
Proof: We prove the theorem by induction on k.
 n + 1
By (3) of property 1.2.42, we get   = n! (1)
 1 
Also since, n! ≡ 0 (mod p), for n ≥ p (2)
Therefore from (1) and (2) we get ,
 n + 1
 1  ≡ 0 (mod p), for n ≥ p
 
n
(or)   ≡ 0 (mod p ), for n -1 ≥ p
1 
n
i.e.,   ≡ 0 (mod p ), for n > p
1 
Therefore the theorem is true for k = 1 .
Assume t he theorem is true for k = m ,
n
That is,  m  ≡ 0 (mod p ) if n > mp
 
(3)
Taking k = m + 1 in theorem 3.4.1 we obtain,
n + p   n  n
 m + 1 ≡  m + 1 − p  + ( p − 1)  m  (mod p) (4)
     
By using (3) in (4) we get,
n + p   n 
 m + 1 ≡  m + 1 − p  (mod p ) if n > mp (5)
   
Since, n > mp > mp + p − p 2 = (m + 1 − p ) p , so using (3) in (5) we get,

n + p 
 m + 1 ≡ 0 (mod p) if n > mp
 
n + p 
(or)   ≡ 0 (mod p ) if n + p > ( m + 1) p
 m + 1
 n 
i.e.,   ≡ 0 (mod p ) if n > ( m + 1) p .
 m + 1
Which shows that the theorem is true for k = m + 1 .
Hence by induction, we obtain
 n
 k  ≡ 0 (mod p ), ∀ n > kp .
 

CHAPTER IV

SOME CONGRUENCES ON BINOMIAL


COEFFICIENTS AND STIRLING NUMBERS OF
BOTH KINDS IN THE CASE G ═ Cp × Cp

4.1 Möbius function defined on L (Cp × Cp):

In this section, we find Möbius function defined on L(CP × CP ) , and hence we


prove a congruence relation from corollary 2.2.7.

Lemma 4.1.1: The Möbius function µ defined on L(CP × CP ) is given by,


 1 if H = 1

µ ( H ) = −1 if H = p

p if H = p 2

Proof: Let G = CP × CP acts on a set S. For notational convenience let σ , τ denote the

permutations given by σ = (1, 2,..., p ) , τ = ( p + 1, p + 2,..., p + p) and for A ⊆ G let A

denotes the sub group generated by A. Let L (C P × C P ) =

{ { e} , σ }
, τ , σ .τ , σ .τ 2 ,..., σ .τ p −2 , σ .τ p −1 , σ ,τ = G . Then ( L(CP × CP ), ⊆ ) is a

lattice and its lattice diagram is,

If H = { e} , then H = 1 . Since { e} is the unique minimal element in L(CP × CP ) ,


then by definition 2.2.1 we get,
µ ({ e} ) = 1 , i.e., µ ( H ) = 1 if H = 1 .

If H = σ , then H = σ = p , and hence,


µ (H ) = µ ( σ ) = − ∑
H1 < σ
µ ( H1 )
H1∈L ( CP ×C P )

= − µ ({ e} )

= −1 .

If H = τ , then H = τ = p , and hence,

µ (H ) = µ ( τ ) = − ∑
H1 < τ
µ ( H1 )
H1∈L (C P ×CP )

= − µ ({ e} )

= −1 .

If H = σ .τ , for all i = 1, 2,..., p − 1 , then H = σ .τ = p , and hence,


i i

µ ( H ) = µ ( σ .τ i ) = − ∑ i
µ ( H1 )
H1 < σ .τ
H1∈L ( CP ×C P )

= − µ ({ e} )

= −1 , for all i = 1, 2,..., p − 1 .

If H = σ ,τ = G = CP × CP , then H = CP × CP = p , and hence,


2

µ ( H ) = µ (CP × CP ) = − ∑
H1 <CP ×CP
µ ( H1 )
H1∈L ( CP ×CP )

{ }
= − µ ( σ ) + µ ( τ ) + µ ( σ .τ ) + µ ( σ .τ 2 ) + ... + µ ( σ .τ p −1 ) + µ ({ e} )

 
= − (−1) + (−1) + ... + (−1) + (−1) + 1
1 44 2 4 43
 ( p −1) times 

= −( − p )
=p
Hence, the Möbius function for each H ≤ G ,
 1 if H = 1

µ ( H ) = −1 if H = p

p if H = p 2
Theorem 4.1.2: If G = CP × CP is a group which acts on a finite set S, then
p −1
β ( { e} ) + pβ ( G ) ≡ β ( σ ) +∑β ( σ .τ i ) + β ( τ ) ( mod p ) . 2

i =1

Proof: By corollary 2.2.7, we get


H ∈L (C P ×C P )
µ ( H ) β ( H ) ≡ 0 (mod CP × CP ) .

That is,

p −1
µ ({ e} ) β ( { e} ) + µ ( σ ) β ( σ ) + ∑ µ ( σ .τ
i =1
i
(
) β σ .τ i ) + µ ( τ )β ( τ ) + µ (G)β ( G ) ≡ 0 ( mod p )
2

Using lemma 4.1.1, we get


p −1
β ( { e} ) − β ( σ ) −∑β ( σ .τ i ) − β ( τ ) + pβ ( G ) ≡ 0 ( mod p ) 2

i =1

p −1

i.e., β ( { e} ) + pβ ( G ) ≡ β ( σ ) +∑β ( σ .τ i ) + β ( τ ) ( mod p ) . 2

i =1

4.2. Congruence relation on Binomial coefficients in the case


G ═ Cp × Cp :

In this section, we prove congruence relation [] on binomial coefficients in the

case G = CP × CP by using theorem 4.1.2.

Theorem 4.2.1: Let S be the set of all k element sub sets T of [ n + 2 p ] and let

G = CP × CP acts on S in a natural way. Then

n + 2p  n + p   n + p    n   n   n  
 − 2  +   +   + 2  +   ≡ 0 ( mod p )
2

 k   k   k − p    k   k − p   k − 2 p  
Proof: Let S = { T : T = { t1 , t2 ,..., tk } ⊆ [ n + 2 p ] } . Then by the definition of binomial

coefficients,
n+ 2p
S = 
 k 
(1)
Let G = CP × CP acts on S in natural way. If every k-sub set T of [ n + 2 p ] is stabilized by

{ e} , that is GT = { e} , ∀T ∈ S , then by theorem 2.1.17, definition 2.1.13and (1) we


obtain,

β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP ×C P )

= ∑ {T ∈S :G
K ≥{ e}
T = K}
K ∈L ( CP ×C P )

= { T ∈ S : GT = { e} } + { T ∈ S : GT = σ } +
p −1

∑ {T ∈S :G
i =1
T = σ .τ i } + {T ∈S :G
T = τ }
= S + ∅ + ... + ∅ , since GT = { e} , ∀T ∈ S

= S

n+ 2p
=  (2)
 k 

Let σ = (1, 2,..., p ) , τ = ( p + 1, p + 2,..., p + p) . Let A denote the sub group


generated by A.
If GT = σ , ∀T ∈ S , then we get two cases.

If 1 ∈ T , then the repeated action of (1, 2,..., p) forces [ p] ⊆ T , that is

{ 1, 2,..., p} ⊆ T = { t1 , t2 ,..., tk } . Now we must choose the k − p elements of T from the set

p + [ n + p ] = { p + 1. p + 2,..., p + p, 2 p + 1, 2 p + 2,..., 2 p + n} .
 n + p
Thus we have   choices for T.
k − p 

Hence, { T ∈ S : GT = σ ,1 ∈ T }

{
= T ∈ S : T = { 1, 2,..., p, t1 , t2 ,..., tk − p } and { t1 , t2 ,..., tk − p } ⊆ p + [ n + p ] }
 n + p
=  (3)
k − p 

If 1 ∉ T , then we have T ⊆ p + [ n + p ] ,

i.e., T = { t1 , t2 ,..., tk } ⊆ { p + 1. p + 2,..., p + p, 2 p + 1, 2 p + 2,..., 2 p + p} .

n + p
Thus we have   choices for T.
 k 

Hence, { T ∈ S : GT = σ ,1 ∉ T } = { T ∈ S : T = { t1 , t2 ,..., tk } ⊆ p + [ n + p ] }
n+ p
=  (4)
 k 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,

β( σ )= ∑ α (K )
σ K≥
K ∈L ( CP ×C P )

= ∑ {T ∈S :G
K≥ σ
T = K}
K ∈L ( CP ×CP )

∑ {T ∈S :G T = K} + ∑ {T ∈S :G T = K}
= K≥ σ K≥ σ
K ∈L ( CP ×CP ) K ∈L ( CP ×CP )
1∈T 1∉T

= { T ∈ S : GT = σ ,1 ∈ T } + { T ∈ S : GT = σ ,1∉ T } +
{ T ∈ S : GT = CP × CP ,1∈ T } + { T ∈ S : GT = CP × CP ,1 ∉ T }

 n + p n + p
= + , since, GT = σ , ∀T ∈ S
k − p   k 
(5)
Suppose, GT = τ , ∀T ∈ S , again we get two possibilities.

If p + 1∈ T , then the repeated action of ( p + 1, p + 2,..., p + p ) forces p + [ p ] ⊆ T ,

that is { p + 1, p + 2,..., p + p} ⊆ T = { t1 , t2 ,..., tk } . Now we must chose the k − p elements

of T from the set { 1.2,..., p, 2 p + 1, 2 p + 2,..., 2 p + n} .

 n + p
Thus we have   choices for T.
k − p 
 n + p
Hence, { T ∈ S : GT = τ , p + 1∈ T } =   (6)
k − p 

If p + 1∉ T , then we have T = { t1 , t2 ,..., tk } ⊆ { 1.2,..., p, 2 p + 1, 2 p + 2,..., 2 p + n} , .

n + p
Thus we have   choices for T.
 k 
n + p
Hence, { T ∈ S : GT = τ ,1 ∉ T } =   (7)
 k 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,

β( τ )= ∑ α (K )
τ K≥
K ∈L (C P ×C P )

= ∑ {T ∈S :G
K≥ τ
T = K}
K ∈L ( CP ×CP )

∑ {T ∈S :G T = K} + ∑ {T ∈S :G T = K}
= K≥ τ K≥ τ
K ∈L ( CP ×CP ) K ∈L ( CP ×CP )
1∈T 1∉T
= { T ∈ S : GT = τ ,1 ∈ T } + { T ∈ S : GT = τ ,1 ∉ T } +

{ T ∈ S : GT = CP × CP ,1∈ T } + { T ∈ S : GT = CP × CP ,1 ∉ T }

 n + p n + p
= + , since GT = τ , ∀T ∈ S
k − p   k 
(8)
Suppose GT = σ .τ , ∀T ∈ S , we have four possibilities depending on which of

the two elements 1, p + 1 are in T.


If 1 ∈ T and p + 1∈ T , then 2,3,…,p and p + 2, p + 3,..., p + p are also in T, that is

{ 1, 2,..., p, p + 1, p + 2,..., p + p} ⊆ T = { t1 , t2 ,..., tk } . Now we choose the remaining k − 2 p

 n 
elements of T from the set 2 p + [ n ] . Thus we have   ways for T.
k − 2p
 n 
Therefore, { T ∈ S : GT = σ .τ , 1∈ T , p + 1 ∈ T } =   (9)
k − 2p

If 1∉ T and p + 1∈ T , then p + 2, p + 3,..., p + p are also in T, that is

{ p + 1, p + 2,..., p + p} ⊆ T ⊆ 2 p + [ n] , that is T = { p + 1, p + 2,..., p + p, t1 , t2 ,..., tk − p }

 n 
⊆ 2 p + [ n ] . Thus we have   ways for T.
k − p
 n 
Therefore, { T ∈ S : GT = σ .τ , 1∉ T , p + 1 ∈ T } =   (10)
k − p
If 1∈ T and p + 1∉ T , then 2,3,..., p are also in T, that is

{ 1, 2,..., p} ⊆ T ⊆ 2 p + [ n ] , that is T = { 1, 2,..., p, t1 , t2 ,..., tk − p } ⊆ 2 p + [ n] . Thus we have

 n 
  ways for T.
k − p
 n 
Therefore, { T ∈ S : GT = σ .τ , 1∈ T , p + 1 ∉ T } =   (11)
k − p

If 1 ∉ T and p + 1∉ T , then we have T = { t1 , t2 ,..., tk } ⊆ 2 p + [ n ] . Thus we have

n
  ways for T.
k 
n
Therefore, { T ∈ S : GT = σ .τ , 1∉ T , p + 1 ∉ T } =   (12)
k 
In view of theorem 2.1.17, definition 2.1.13 and equations (9), (10), (11) and (12) we get,
β ( σ .τ )= ∑ α (K )
στ K≥ .
K ∈L ( CP ×CP )

= ∑ {T ∈S :G
K ≥ σ .τ
T = K}
K ∈L ( CP ×C P )

= { T ∈ S : GT = σ .τ ,1 ∈ T , p + 1∈ T }

+ { T ∈ S : GT = σ .τ ,1∉ T , p + 1 ∈ T }

+ { T ∈ S : GT = σ .τ ,1∈ T , p + 1 ∉ T }

+ { T ∈ S : GT = σ .τ ,1∉ T , p + 1 ∉ T }

+ { T ∈ S : GT = CP × CP ,1 ∈ T , p + 1∈ T }

+ { T ∈ S : GT = CP × CP ,1 ∉ T , p + 1∈ T }

+ { T ∈ S : GT = CP × CP ,1 ∈ T , p + 1∉ T }

+ { T ∈ S : GT = CP × CP ,1 ∉ T , p + 1∉ T }
 n   n   n  n
= + +  +   , since GT = σ .τ , ∀T ∈ S
k −2p k − p k − p k 
 n   n  n
=  + 2 + 
k −2p k − p k 
(13)

Similarly if GT = σ .τ , ∀T ∈ S , 2 ≤ i ≤ p − 1 , then we get


i

) =  k − 2 p  + 2  k − p  +  k  , for i = 2,3,..., p − 1


n n n
(
β σ .τ i (14)

And if GT = σ ,τ = CP × CP , ∀T ∈ S , then we get

 n   n  n
β ( G) =   + 2 +  (15)
k − 2p k − p k 
Hence in view of theorem 4.1.2, equations (2), (5) , (8), (13), (14) and (15) we get
 n   n   n    n + p   n + p    n   n   n  
 + 2  +    ≡ 2  +   + ( p − 1)   + 2  +    ( mod p )
2
p 
 k − 2 p   k − p   k    k − p   k    k − 2 p   k − p   k  

n + 2p  n + p   n + p     n   n   n  
 − 2  +   +   + 2  +   ≡ 0 ( mod p ) .
2
i.e., 
 k   k − p   k     k   k − p   k − 2 p  

4.3. Congruence relation on Stirling numbers of the second


kind in the case G ═ Cp × Cp :
In this section, we prove congruence relation [] on Stirling numbers of the second

kind in the case G = CP × CP by using theorem 4.1.2.

Theorem 4.2.1: Let S be the set of all partitions π of [ n + 2 p ] and let G = CP × CP acts
on S in a natural way. Then
n + 2 p  1
 n + p + i  2  2  n + i   n 
 − 2∑   + ∑    − p ( p − 1)   ≡ 0 ( mod p )
2

 k  i = 0 k + (i − 1) p  i =0  i  k + (i − 2) p  k − p 

Proof: Let S = { π : π = { T1 , T2 ,..., Tk } , Ti ⊆ [ n + 2 p ] ,1 ≤ i ≤ k } and by the definition of

Stirling numbers of second kind, we have


n + 2 p 
S = 
 k 
(1)
Let G = CP × CP acts on S in natural way,

If Gπ = { e} , ∀π ∈ S , then by theorem 2.1.17, definition 2.1.13and (1) we obtain,

β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP ×CP )

= ∑ {π ∈S :G
K ≥{ e}
π = K}
K ∈L ( CP ×C P )

= { π ∈ S : Gπ = { e} } + { π ∈ S : Gπ = σ } +
p −1

∑ {π ∈S :G
i =1
π = σ .τ i } + {π ∈S :G
π = τ }
= S + ∅ + ... + ∅ , since Gπ = { e} , ∀π ∈ S

= S

n + 2 p 
=  (2)
 k 
Let σ = (1, 2,..., p ) , τ = ( p + 1, p + 2,..., p + p) . Let A denote the sub group
generated by A.
If Gπ = σ , ∀π ∈ S , then we get two possibilities.

If { 1} is a singleton block of π , then so are { 2} , { 3} ,..., { p} . That is

{ { 1} , { 2} , { 3} ,...,{ p} } ⊆ π = { T , T ,..., T } ,
1 2 k where Ti ⊆ [ n + 2 p ] ,1 ≤ i ≤ k . Now, we

partitioning the remaining k − p blocks of π from the set p + [ n + p ] . This means there

n + p 
are   ways to chose the rest of the blocks of π .
k − p 

Thus, { π ∈ S : Gπ = σ , { 1} ∈ π}

{
= π ∈ S : π = { { 1} , { 2} , { 3} ,..., { p} , T1 , T2 ,..., Tk − p } and { T1 , T2 ,..., Tk − p } ⊆ p + [ n + p ] }
n + p 
=  (3)
k − p 

If { 1} ∉ π , then 1 is in a block of size at least two, so 2,3,….,p are also in the same

block. Thus the set [ p ] = { 1, 2,..., p} is acting like a single element, which we shall
denote by P, so we are really just partitioning the set

{ P, p + 1, p + 2,..., p + p, 2 p + 1, 2 p + 2,..., 2 p + n} into k-blocks. This means there are

 n + p + 1
  ways to choose the blocks of π .
 k 

Thus, { π ∈ S : Gπ = σ , { 1} ∉ π}

= { π ∈ S : π = { T1 , T2 ,..., Tk } and Ti ⊆ { P, p + 1, p + 2,..., p + n} ,1 ≤ i ≤ k }

 n + p + 1
=  (4)
 k 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,
β( σ )= ∑ α (K )
σ K≥
K ∈L ( CP ×C P )

=
K≥ σ
∑ { π∈ S :G π = K}
K ∈L ( CP ×C P )

= ∑ { π∈ S :G
K≥ σ
π = K} + ∑ { π∈ S :G
K≥ σ
π = K}
K ∈L ( CP ×C P ) K ∈L ( CP ×CP )
{ 1} ∈π {1}∉π

= { π ∈ S : Gπ = σ , { 1} ∈ π} + { π ∈ S : Gπ = σ , { 1} ∉ π}

+ { π ∈ S : Gπ = CP × CP , { 1} ∈ π} + { π ∈ S : Gπ = CP × CP , { 1} ∉ π}

 n + p   n + p + 1
= +   , since Gπ = σ , ∀π ∈ S (5)
k − p   k 

If Gπ = τ , ∀π ∈ S , then we get two possibilities.

If { p + 1} is a singleton block of π , then so are { p + 2} , { p + 3} ,..., { p + p} . That

is { { p + 1} ,{ p + 2} ,...,{ p + p} } ⊆ π = { T , T ,..., T } , where


1 2 k Ti ⊆ [ n + 2 p ] ,1 ≤ i ≤ k . Now,

we partitioning the remaining k−p blocks of π from the set

n + p 
{ 1, 2,..., p, 2 p + 1, 2 p + 2,..., 2 p + n} . This means there are   ways to chose the rest
k − p 
of the blocks of π .
n + p 
Thus, { π ∈ S : Gπ = τ , { p + 1} ∈ π} =   (6)
k − p 

If { p + 1} ∉ π , then 1 is in a block of size at least two, so p + 2, p + 3,..., p + p are

also in the same block. Thus the set [ p + 1] = { p + 1, p + 2,..., p + p} is acting like a single
element, which we shall denote by Q, so we are really just partitioning the set

 n + p + 1
{ 1, 2,..., p, Q, 2 p + 1, 2 p + 2,..., 2 p + n} into k-blocks. This means there are  
 k 
ways to choose the blocks of π .
 n + p + 1
Thus { π ∈ S : Gπ = τ , { p + 1} ∉ π} =   (7)
 k 
Hence by theorem 2.1.17, definition 2.1.13 and equations (6) and (7) we get,
β( τ )= ∑ α (K )
τ K≥
K ∈L (C P ×C P )

=
K≥ τ
∑ { π∈ S :G π = K}
K ∈L ( CP ×CP )

= ∑ { π∈ S :G
K≥ τ
π = K} + ∑ { π∈ S :G
K≥ τ
π = K}
K ∈L ( CP ×CP ) K ∈L ( CP ×CP )
{ p +1} ∈π { p +1}∉π

= { π ∈ S : Gπ = τ , { p + 1} ∈ π} + { π ∈ S : Gπ = τ , { p + 1} ∉ π}

+ { π ∈ S : Gπ = CP × CP , { p + 1} ∈ π} + { π ∈ S : Gπ = CP × CP , { p + 1} ∉ π}

 n + p   n + p + 1
= +   , since Gπ = τ , ∀π ∈ S (8)
k − p   k 

Suppose Gπ = σ .τ , ∀π ∈ S , then there are five possibilities.

If the sets [ p] and p + [ p ] are divided into singleton blocks, then we have

{ { 1} , { 2} ,..., { p} , { p + 1} , { p + 2} ,..., { p + p} } ⊆ π , where π = { T1 , T2 ,..., Tk } and

Ti ⊆ [ n + 2 p ] , 1 ≤ i ≤ k . Now we partition the remaining k − 2 p blocks of π from the

 n 
set 2 p + [ n ] . Thus there are   ways to choose the rest of the blocks of π .
k − 2 p 

If the set [ p ] is one block and p + [ p ] divided into singleton blocks, then the set

[ p ] = { 1, 2,..., p} is acts like a single element and we shall dente this by P, and

{ p + 1} , { p + 2} ,..., { p + p} are singleton blocks of π, that is

π = { { p + 1} , { p + 2} ,..., { p + p} , T1 , T2 ,..., Tk − p } . Now we partition the set


 n +1 
{ P, 2 p + 1, 2 p + 2,..., 2 p + n} into k − p blocks of π . Thus there are   ways to
k − p 
choose the rest of the blocks of π .

If the set [ p ] is divided into singleton blocks and p + [ p ] is one block, then the

set p + [ p ] = { p + 1, p + 2,..., p + p} acts like a single element and we shall denote this by

Q , and also we have { 1} , { 2} , { 3} ,..., { p} are singleton blocks of π , that is

π = { { 1} , { 2} ,..., { p} , T1 , T2 ,..., Tk − p } . Now we partition the set

 n +1 
{ Q, 2 p + 1, 2 p + 2,..., 2 p + n} into remaining k − p blocks of π . Thus there are  
k − p 
ways to choose the rest of the blocks of π .
If both the sets [ p ] and p + [ p ] are single blocks, then both of them are acting
like single elements, which we shall denote by P and Q respectively. Therefore we

partition the set { P, Q, 2 p + 1, 2 p + 2,..., 2 p + n} into k blocks. This means there are

 n + 2
  ways to choose the blocks of π .
 k 

Suppose there are partitions stabilized by σ .τ consisting of p doubletons of the

form { 1, t} , { 2,τ (t )} ,..., { p,τ (t )} , where t ∈ p + [ p ] , and a partition of 2 p + [ n ] into


p −1

 n 
k − p parts, for which there are p   possibilities.
k − p 

Hence, in view of theorem 2.1.17, definition 2.1.13 and above five possibilities
we get,
β ( σ .τ )= ∑ α (K )
στ K≥ .
K ∈L ( CP ×CP )
= ∑ {π ∈S :G
K ≥ σ .τ
π = K}
K ∈L ( CP ×C P )

 n   n + 1   n + 1   n + 2  n 
= + +  + + p ,
k − 2 p   k − p  k − p   k  k − p 

since Gπ = σ .τ , ∀π ∈ S

 n   n + 1   n + 2  n 
= + 2 +  + p  (9)
k − 2 p  k − p   k  k − p 

Similarly if GT = σ .τ , ∀T ∈ S , 2 ≤ i ≤ p − 1 , then we get


i

n +1 n + 2
) = k − 2 p  + 2 k − p  +   n 
n
(
β σ .τ i
k 
+ p 
k − p 
(10)

And if G = σ ,τ = CP × CP acting on partitions, where the sets [ p ] and p + [ p ] are


either divided into singletons or are contained wholly in one block. Then we get
 n   n + 1   n + 2
β ( G) =  + 2 + 
k − 2 p  k − p   k 
(11)
Hence in view of theorem 4.1.2, equations (2), (5) , (8), (9), (10) and (11) we get
p −1
β ( { e} ) + pβ ( G ) − β ( σ ) −β( τ ) − +∑ β ( σ .τ i ) + β ( τ ) ≡ 0 ( mod p )
2

i =1

n + 2 p   n   n + 1   n + 2    n + p   n + p + 1 
i.e.,  + p + 2 +   − 2  + 
 k    k − 2 p   k − p   k     k − p   k 

 n   n + 1   n + 2  n 
−( p − 1)   + 2 + + p   ≡ 0 ( mod p )
2

 k − 2 p  k − p   k  k − p  

n + 2 p  1
 n + p + i  2  2  n + i   n 
 − 2∑   + ∑    − p ( p − 1)   ≡ 0 ( mod p )
2
i.e., 
 k  i = 0 k + (i − 1) p  i =0  i  k + (i − 2) p  k − p 

Following theorem gives a congruence relation bell numbers.

Theorem 4.3.2
1 2
 2
B ( n + 2 p ) − 2∑ B ( n + p + i ) + ∑   B ( n + i ) − p ( p − 1) ≡ 0 ( mod p 2 )
i =0 i =0  i 

Proof: By summing the result in theorem 4.3.1, for all values of k, we get
n + 2 p  1
 n+ p+i  2
 2  n + i   n 
∑  − 2∑∑   + ∑∑     − p( p − 1)∑   ≡ 0 ( mod p )
2

k  k  k i = 0  k + (i − 1) p  k i = 0  i   k + (i − 2) p  k k − p 

i.e.,

n + 2 p  1
 n+ p+i  2  2  n + i   n 
∑k  k  ∑∑ −  + ∑∑     − p ( p − 1)∑   ≡ 0 ( mod p )
2
2 
  i =0 k k + (i − 1) p  i =0 k  i  k + (i − 2) p  k k − p 

1 2
 2
i.e., B ( n + 2 p ) − 2∑ B ( n + p + i ) + ∑   B ( n + i ) − p ( p − 1) ≡ 0 ( mod p ) .
2

i =0 i =0  i 

4.4. Congruence relation on Stirling numbers of the first kind


in the case G ═ Cp × Cp :
In this section, we prove congruence relation [] on Stirling numbers of the first

kind in the case G = CP × CP by using theorem 4.1.2.


Theorem 4.2.1: Let S be the set of all permutations σ ′ with k disjoint cycles defined on

[ n + 2 p] and let G = CP × CP acts on S by conjugation. Then

n + 2 p  1
 n+ p  2
 2  n   n 
 k  − 2∑  k + (i − 1) p − i  ( p − 1) + ∑  i   k + (i − 2) p − i  − p ( p − 1)  k − p  ≡ 0 ( mod p )
i 2

  i=0   i =0      

Proof: Let S = { σ : σ is a permutation of [ n + 2 p ] with k disjoint cycles} and by the


′ ′

definition of Stirling numbers of first kind, we have


n + 2 p 
S =  (1)
 k 
Let G = CP × CP acts on S by conjugation.

If every permutation σ′ of [ n + 2 p ] is stabilized by { e} , that is Gσ′ = { e} , ∀σ ∈ S , then


by theorem 2.1.17, definition 2.1.13and (1) we obtain,


β ( { e} ) = ∑ α (K )
K ≥{ e}
K ∈L ( CP ×CP )

= ∑ { σ ∈S :G
K ≥{ e}

σ′
=K }
K ∈L ( CP ×CP )

{ } {
= σ′ ∈ S : Gσ′ = { e} + σ′ ∈ S : Gσ′ = σ }
p −1

i =1
{
+ ∑ σ′ ∈ S : Gσ′ = σ .τ i } + { σ ∈S :G

σ′
= τ }
{
+ σ′ ∈ S : Gσ′ = CP × CP }
= S + ∅ + ∅ + ... + ∅ , since Gσ′ = { e} , ∀σ′ ∈ S

n + 2 p 
=  (2)
 k 

Suppose Gσ′ = σ , ∀σ ∈ S , then to compute β ( σ ) there are two possibilities.


If 1 is in a cycle τ ′ of length 1, then ( 1) is the one cycle of the product of cycles

of σ′ and so ( 2 ) , ( 3) ,..., ( p ) are also cycles of the product of cycles σ′ . Now, we must

choose the remaining k − p cycles of σ′ from the set p + [ n + p ] . This means there are

n + p  ′
 k − p  ways to chose the rest of the cycles of σ .
 

{
Thus, σ ∈ S : Gσ′ = σ ,1∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S
′ ′ ′ ′ ′
}
= { σ ∈ S : σ = ( 1) ( 2 ) ... ( p ) τ 1τ 2 ...τ k − p and τ i is a cycle on p + [ n + p ] ,1 ≤ i ≤ k − p }
′ ′ 1 1 1 1

n + p 
=   (3)
k − p 

If 1 is in a cycle τ ′ of length at least two, then 2,3,….,p are also in the same

cycle, i.e., τ ′ = (1, 2,3,..., p ) . We chose remaining k − 1 cycles of σ′ from the set
n + p 
p + [ n + p ] . This means there are  ′
 ways to construct the rest of the cycles of σ ,
 k − 1 

and also there are p − 1 choices for τ ′ , i.e., τ ′ = (1, 2,3,..., p) j , for 1 ≤ j < p . Therefore

n + p 
there are ( p − 1)  ′
 choices for cycles of σ .
 k − 1 

Thus, { σ ∈ S : Gσ = σ ,1∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }


′ ′ ′ ′ ′

= { σ ∈ S : σ = τ τ 1τ 2 ...τ k − p and τ i is a cycle on p + [ n + p ] ,1 ≤ i ≤ k − p }


′ ′ ' ' ' ' '

n + p 
= ( p − 1)   (4)
k − 1 
In view of theorem 2.1.17, definition 2.1.13 and equations (3) and (4) we get,
β( σ )= ∑ α (K )
σ K≥
K ∈L ( CP ×CP )

= ∑ { σ ∈S :G
K≥ σ

σ′
=K }
K ∈L ( CP ×CP )

{
= σ ∈ S : Gσ′ = CP ,1 ∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S
′ ′ ′ ′ ′
}
+ { σ ∈ S : Gσ = σ ,1∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }
′ ′ ′ ′ ′

n + p  n + p 
, since Gσ′ = σ , ∀σ ∈ S

=  +( p − 1)   (5)
k − p  k − 1 

Suppose Gσ′ = τ , ∀σ ∈ S , then to compute β ( τ ) there are two possibilities.


If p + 1 is in a cycle τ ′′ of length 1, then ( p + 1) is the one cycle of the product of

cycles of σ′ and so ( p + 2 ) , ( p + 3) ,..., ( p + p ) are also cycles of the product of cycles σ′

. Now, we must choose the remaining k − p cycles of σ′ from the set

n + p 
{ 1, 2,3,..., p, 2 p + 1, 2 p + 2,..., 2 p + n} . This means there are   ways to chose the
k − p 
rest of the cycles of σ′ .
{
Thus, σ ∈ S : Gσ′ = τ , p + 1 ∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S
′ ′′ ′′ ′′ ′
}
n + p 
=   (6)
k − p 

If p + 1 is in a cycle τ ′′ of length at least two, then p + 2, p + 3,..., p + p are also

in the same cycle, i.e, τ ′′ = ( p + 1, p + 2, p + 3,..., p + p ) . We chose remaining k − 1 cycles

n + p 
of σ′ from the set { 1, 2,3,..., p, 2 p + 1, 2 p + 2,..., 2 p + n} . This means there are  
k − 1 
ways to construct the rest of the cycles of σ′ , and also there are p − 1 choices for τ ′′ , i.e.,

n + p 
τ ′′ = ( p + 1, p + 2,..., p + p ) j , for 1 ≤ j < p . Therefore there are ( p − 1)   choices for
k − 1 
cycles of σ′ .

Thus, { σ ∈ S : Gσ = τ ,1 ∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }


′ ′′ ′′ ′′ ′

n + p 
= ( p − 1)   (7)
k − 1 
In view of theorem 2.1.17, definition 2.1.13 and equations (6) and (7) we get,
β( τ )= ∑ α (K )
τ K≥
K ∈L (C P ×C P )

= ∑ { σ ∈S :G
K≥ τ

σ′
=K }
K ∈L ( CP ×CP )

{
= σ ∈ S : Gσ′ = τ , p + 1 ∈τ , l (τ ) = 1, where τ is a cycle of σ ∈ S
′ ′′ ′′ ′′ ′
}
+ { σ ∈ S : Gσ = τ ,1 ∈τ , l (τ ) ≥ 2, where τ is a cycle of σ ∈ S }
′ ′′ ′′ ′′ ′

n + p  n + p 
 , since Gσ′ = τ , ∀σ ∈ S

=  +( p − 1)  (8)
 k − p   k − 1 

Suppose Gσ′ = σ.τ , ∀σ ∈ S , then there are five possibilities.



If 1 is in a cycle τ ′ of length 1 and p + 1 is in a cycle τ ′′ of length 1, then

( 1) , ( 2 ) , ( 3) ,..., ( p ) , ( p + 1) , ( p + 2 ) , ( p + 3) ,..., ( p + p ) are cycles of σ′ . We choose the

 n 
remaining k − 2 p cycles of σ′ from the set 2 p + [ n ] . Thus there are   ways to
k − 2 p 
choose the rest of the blocks of σ′ .
If 1 is in a cycle τ ′ of length at least 2 and p + 1 is in a cycle τ ′′ of length 1, then

τ ′ = (1, 2,3,..., p ) , ( p + 2 ) , ( p + 3) ,..., and ( p + p ) are the cycles of σ′ . We choose the

 n 
remaining k − ( p + 1) cycles of σ′ from the set 2 p + [ n ] . Therefore there are  
 k − p − 1

ways to construct rest of the cycles of σ′ . Also we have ( p − 1) choices for τ ′ , given by

 n 
τ ′ = (1, 2,3,..., p) j , for 1 ≤ j < p . Therefore there are ( p − 1)  ′
 choices for σ in
 k − p − 1
this case.

If 1 is in a cycle τ ′ of length 1 and p + 1 is in a cycle τ ′′ of length at least two,

then ( 1) , ( 2 ) , ( 3) ,..., ( p ) and τ ′′ = ( p + 1, p + 2, p + 3,..., p + p ) are cycles of σ′ . Now we

choose the remaining k − ( p + 1) cycles of σ′ from the set 2 p + [ n ] . Therefore there are

 n 
 k − p − 1 ways to construct rest of the cycles of σ . Also we have ( p − 1) choices for

 

τ ′′ , given by τ = ( p + 1, p + 2,..., p + p ) , for 1 ≤ j < p .


′′ j
Therefore there are

 n 
( p − 1)  ′
 choices for σ in this case.
 k − p − 1
If 1 is in a cycle τ ′ of length at least 2 and p + 1 is in a cycle τ ′′ of length at least

two, then τ ′ = (1, 2,3,..., p ) and τ ′′ = ( p + 1, p + 2, p + 3,..., p + p ) are two cycles of σ′ .


Now we choose the remaining k − 2 cycles of σ′ from the set 2 p + [ n ] . Therefore there

 n 
are   ways to construct rest of the cycles of σ′ . Also we have ( p − 1) choices for
k − 2

both τ ′′ and τ ′′ , they are given by τ ′ = (1, 2,3,..., p) j and τ ′′ = ( p + 1, p + 2,..., p + p ) j , for

2  n 
1 ≤ j < p . Therefore there are ( p − 1)  ′
 choices for σ in this case.
 k − 2 

Suppose there are cycles stabilized by σ .τ consisting of p transpositions of the

form ( 1, t ) , ( 2,τ (t ) ) ,..., ( p,τ (t ) ) , t ∈ p + [ p ] . We must choose remaining k − p choices


p −1

 n 
of σ′ from the set 2 p + [ n ] , for which there are p   possibilities.
k − p 

Hence, in view of theorem 2.1.17, definition 2.1.13 and above five possibilities
we get,
β ( σ .τ )= ∑ α (K )
στ K≥ .
K ∈L (C P ×C P )

= ∑ {π ∈S :G
K ≥ σ .τ
π = K}
K ∈L ( CP ×CP )

 n   n   n   n   n 
=  + ( p − 1)   + ( p − 1)   + ( p − 1) 2  + p ,
k − 2 p   k − p − 1  k − p − 1 k − 2 k − p 

since Gσ′ = σ .τ , ∀σ ∈ S

 n   n  2  n   n 
=  + 2( p − 1)  k − p − 1 + ( p − 1) k − 2 + p  (9)
k − 2 p      k − p 

Similarly we get if Gσ′ = σ .τ i , ∀σ′ ∈ S , 2 ≤ i ≤ p − 1 , then we get

) = k − 2 p  + 2( p − 1) k − p − 1 + ( p − 1)  n   n 


n n
(
β σ .τ i 2
k − 2 +
 
p 
k − p 
(10)
And if G = σ ,τ = CP × CP acting on permutations, where the cycles τ ′ = (1, 2,3,..., p )

and τ ′′ = ( p + 1, p + 2, p + 3,..., p + p ) are either divided into disjoint cycles of length 1 or


they are cycles of length p. Then we get
 n   n   n 
β ( G) =   + 2( p − 1)   + ( p − 1) 2  
k − 2 p   k − p − 1 k − 2
(11)
Hence in view of theorem 4.1.2, equations (2), (5) , (8) , (9), and (11) we get
p −1
β ( { e} ) + pβ ( G ) − β ( σ ) −β( τ (
) − +∑ β σ .τ i
i =1
) + β ( τ ) ≡ 0 ( mod p )
2

i.e.,

n + 2 p   n   n   n    n + p  n + p  
 k + p  + 2( p − 1)   + ( p − 1) 2    − 2   +( p − 1)  
   k − 2 p   k − p − 1 k − 2   k − p  k − 1  

 n   n   n   n 
−( p − 1)   + 2( p − 1)  + ( p − 1) 2  +  ≡ 0 ( mod p )
2
  p 
  k − 2 p   k − p − 1  k − 2  k − p  
i.e.,

n + 2 p   n + p  n + p     n   n  2  n 
 k  − 2   k − p  +( p − 1)  k − 1   +   k − 2 p  + 2( p − 1)  k − p − 1 + ( p − 1)  k − 2  
           

 n 
− p ( p − 1)   ≡ 0 ( mod p 2 )
k − p 
i.e.,

n + 2 p  1
 n+ p  2
2  n   n 
 k  ∑  k + (i − 1) p − i 
− 2 ( p − 1) i
+ ∑    − p ( p − 1)   ≡ 0 ( mod p 2 )
  i=0   i = 0  i   k + (i − 2) p − i  k − p 

CHAPTER V
SOME CONGRUENCES ON BINOMIAL
COEFFICIENTS AND STIRLING NUMBERS OF
BOTH KINDS IN THE CASE G =C rp
4.1 Möbius function defined on L ( C p ) :
r

In this section we find the Möbius function defined on L ( C p ) [], and hence we
r

r
prove a congruence relation from corollary 2.2.7. In this section , the group C p is treated

as an r − dimensional vector space over the Galois field GF ( p ) and so the number of

subgroups of order p k is just the number of subspaces of dimension k.


In 1964, G.C. Rota proved the following lemma which calculate the Möbius

function of all the subgroups in L ( C p ) .


r

Lemma 4.1.1. The Möbius function defined on L ( C p ) is given by


r

k ()k
µ ( H ) = ( −1) p 2 , for all subgroups H ≤ C p of order p , where 0 ≤ k ≤ r .
r k

Proof: Since all the subgroups H ≤ C p of order p k are isomorphic to C p , where


r k

r
0 ≤ k ≤ n , so we need only to find the Möbius function of the whole group C p .

That is we prove, µ ( C pr ) = ( −1) p 2


r ( r) (1)

We prove this result by induction on n.


By lemma 4.1.1, we have µ ( { e} ) = 1 ,

µ ( C1p ) = −1 ,

µ ( C p2 ) = p .

Hence the result is true for r ≤ 2 .


Assume the result (1) is true for r − 1 . For the inductive step we choose 1− dimensional

subspace H1 = C p of C p and apply theorem 2.2.4 to get,


1 r

µ ( C pr ) = − ∑ µ( H)
, since { e} < H1 ≤ C p
H <C rp r
H ∨ H1 =C rp
( )
H ∈L C rp
( r2−1)
∑ ( −1)
r −1
=− p
H <C rp
H ’H ,dim( H ) = r −1
H ∈L C rp( )
  r  r − 1   r −1 ( 2 )
r −1
= −  −
   ( − 1) p
  r − 1  r − 2  
  r   r − 1  r −1 ( 2 )
r −1
= −   −   ( )
− 1 p
 1   1  
 p r − 1 p r −1 − 1  r −1 ( 2 )
r −1
= − −  ( −1) p
 p −1 p −1 

= − ( p r −1 ) ( −1)
r −1
p
( r2−1)

( r
= ( −1) p 2 .
r )

Theorem 5.1.3: Let G = C p be a group, which acts on S, and H is a subgroup of G


r

associated with i cycles. Then


r
r
∑ (−1)  i β ( H ) ≡ 0 ( mod p )
i r

i =0  

Proof: Consider the cycles σ j = ( jp + 1, jp + 2,..., jp + p) , 0 ≤ j < r . Given g ∈ C p ,


r

r
g = ∏ σ j j , then σ j is a factor of g if p Œ
n
nj .
j =0

We say that a subgroup a H ≤ C p is associated with i cycles if the set of all factors
r

of elements of H contains exactly i of the cycles σ j . Equivalently H is associated with i

cycles if and only if the smallest subgroup of the form H = σ j1 , σ j 2 ,... containing H

has i generators.
By using corollary 2.2.7 we get,

∑ µ ( H ) β ( H ) ≡ 0 ( mod
H ∈L (G )
C pr ) (1)
r
Since, for each i ( 0 ≤ i ≤ r ), there are   subgroups which are product of i
i 
cycles of the form H , then from (1) we obtain,
r
r
∑  i µ ( H ) β ( H ) ≡ 0 ( mod p ) r
(2)
i =0  
Since H is the smallest subgroup containing H, then it is obvious that
β (H ) = β (H ) (3)
By using (2) and (3) in (1) we obtain,
 
 
r
 r   β ( H ) ≡ 0 ( mod p r )
∑   ∑
i =0  i 

H <H
µ ( H ) 
 H ∈L (C rp ) 
 
 
 
r
 r 
or ∑   ∑ µ ( H )  β ( H ) ≡ 0 ( mod p )
r
(4)
i
i =0    H <H 
 H ∈L (C rp ) 
 
To complete the proof we need only to show that for each choice of H ,


H <H
µ ( H ) = (−1)i
H ∈L (C rp )

By the principle of inclusion and exclusion[ ], the number of subgroups H having

order p k and associated with i cycles is given by,

i   j  j
∑ (−1) i− j
    , where  k  is the q − binomial coefficient for each j (5)
j  j  k  p  p

Since each of these subgroups has Möbius function (−1) k p ( 2 ) , and by using (5)
k

we get,
( k) i   j
∑ µ ( H ) = ∑ ( −1) k p 2 ∑ (−1)
i− j
  
H <H k j  j  k  p
H ∈L (C rp )
i  ( k)
= ∑ ( −1)i − j   ∑ (−1)k p 2
 j
 
j  j k k  p

i 
= ∑ ( −1)i − j   δ j 0
j  j
= (−1)i (6)
Hence by using (6) in (4) we get,
r
r
∑ (−1)  i β ( H ) ≡ 0 ( mod p )
i r

i =0  

5.2. Congruence relation on Binomial coefficients in the case


G =C :rp
In this section, we prove congruence relation [] on binomial coefficients in the

case G = C p by using theorem 5.1.2.


r

Theorem 5.2.1. Let S be the set of all k element subsets of T of [ n + rp ] and let G = C p
r

acts on S in natural way. Then


r
r r
 i  n − ip 
∑ (−1)  i ∑  j  k − jp  ≡ 0 ( mod
i
pr )
i =0   i =0   

Proof: If H is associated with i cycles and stabilizes a subset T ⊆ [ n + rp ] , then each of


the i cycles is either completely contained in or disjoint from T.
i 
Choosing j of the cycles to be in T can be done in   ways and the rest of T can
 j

 n − ip 
be completed in   ways.
 k − jp 
i
 i  n − ip 
Thus, β ( H ) = ∑    (1)
j = 0  j  k − jp 

By theorem 4.1.3 we have,


r
r
∑ (−1)  i β ( H ) ≡ 0 ( mod p )
i r
(2)
i =0  
From (1) and (2) we get,
r
r i
 i  n − ip 
∑ (−1)  i ∑  j  k − jp  ≡ 0 ( mod p )
i r

i =0   j = 0   

5.2. Congruence relation on Stirling numbers of the second


kind in the case G =C :p
r

In this section, we prove congruence relation [] on Stirling numbers of the

second kind in the case G = C p by using theorem 5.1.2.


r

Theorem 5.2.1: Let S be the set of all partitions π of [ n + rp ] into k distinct subsets. Let

G = C pr acts on S. Then,
r
 r  i i  j
 j − 1 + δ ij  n + (r − i ) p + l 
∑ (−1)  ∑   [ − p ( p − 1) ] ∑   ≡ 0 ( mod p )
i i− j r
 (1)
i =0  i  j =0  j  l =0  l   k − ( j − l) p 
Proof: If a subgroup H is associated with i cycles and stabilizes a partition π of

[ n + rp ] , then π induces a partition π of the cycles associated with H. Specifically, π a

and π b are in the same block of π if and only if some element of π a appears in the same

block with some element of π b in π (hence every element of π a appears in a block

i 
with some element of π b in π ). This accounts for   terms in (1) and rest of π can
 j
be completed in
j
 j − 1 + δ ij  n + (r − i ) p + l 
[ − p( p − 1)] ∑ 
i− j
  ways.
l =0  l   k − ( j − l) p 
i
i  j
 j − 1 + δ ij   n + (r − i ) p + l 
Thus, β ( H ) = ∑   [ − p ( p − 1) ] ∑ 
i− j
  (2)
j =0  j  l =0  l   k − ( j − l) p 
By theorem 5.1.3 we have,
r
r
∑ (−1)  i β ( H ) ≡ 0 ( mod p )
i r
(3)
i =0  
From (2) and (3) we get,
r
i r
i
i  j
 j − 1 + δ ij  n + (r − i ) p + l 
∑  ∑ [ ] ∑  ≡ 0 ( mod p )
i− j
( −1)   − p ( p − 1)  
r

i =0 i j
  j =0   l =0  l  k − ( j − l ) p 

5.3. Congruence relation on Stirling numbers of the first kind


in the case G =C :rp
In this section, we prove congruence relation [] on Stirling numbers of the

first kind in the case G = C p by using theorem 5.1.2.


r

Theorem 5.3.1: Let S be the set of all permutations σ of [ n + rp ] into k disjoint cycles.

Let G = C p acts on S by conjugation. Then,


r

r
i r
i
i  j
 j − 1 + δ ij   n + (r − i ) p 
∑  ∑   [ ] ∑  k − ( j − l ) p − l  ≡ 0 ( mod p )
i− j
( −1) − p ( p − 1)   ( p − 1)
l r
(1)
i =0 i j
  j =0   l =0  l   
Proof: If a subgroup H is associated with i cycles and stabilizes a permutation σ of

[ n + rp ] , then σ induces a permutation σ of the cycles associated with H. Specifically,

σ a and σ b are in the same permutation σ if and only if some element of σ a appears in

the same permutation with some element of σ b in σ (hence every element of σ a

i 
appears in a permutation with some element of σ b in σ ). This accounts for   terms
 j
in (1) and rest of σ can be completed in
j
 j − 1 + δ ij  l  n + (r − i ) p 
[ − p( p − 1)] ∑ 
i− j
( p − 1)   ways.
l =0  l  k − ( j − l ) p − l 
i
i  j
 j − 1 + δ ij  l  n + (r − i ) p 
Thus, ( ) ∑   [ ] ∑
i− j
β H = − p ( p − 1) ( p − 1)   (2)
j =0  j  l =0  l  k − ( j − l ) p − l 
By theorem 5.1.3 we have,
r
r
∑ (−1)  i β ( H ) ≡ 0 ( mod p )
i r
(3)
i =0  
From (2) and (3) we get,
r
i r
i
i  j
 j − 1 + δ ij   n + (r − i ) p 
∑  ∑   [ ] ∑  k − ( j − l ) p − l  ≡ 0 ( mod p )
i− j
( −1) − p ( p − 1)   ( p − 1)
l r

i =0 i j
  j =0   l =0  l   

REFERENCES

[1]. G. E. Andrews, The theory of Partitions, Vol. 2, Encyclopedia of


Mathematics and its applications, Addison – Wesley, Reading,
Mass., 1976.

[2]. M. Abramowitz, and , I. A. Stegun (Eds.). Handbook of


Mathematical Functions with Formulas, Graphs, and Mathematical
Tables, 9th printing. New York: Dover, 1972.

[3]. H. W. Becker and J. Riordan, The arithmetic of Bell and Stirling


numbers, Amer. J. math. 70 (1948), 385 - 394.

[4]. L. Carlitz, Weighted Stirling numbers of the first kind and second
kind, I, Fibonacci Quart. 18 (1970), 147 - 162.

[5] L. Carlitz, q-Bernoulli Numbers and Polynomials, Duke Math. J.,


15(1948): 987-1000.

[6]. L. Comtet, Analyse Combinatoire, Tomes i, ii, Presses Univ. de


France, Paris, 1970.

[7]. L. Comtet, Advanced Combinatorics: The Art of Finite and Infinite


Expansions, rev. enl. ed. Dordrecht, Netherlands: Reidel, 1974.

[8]. B. A. Davey and H. A. Priestley, Introduction to Lattices and


Order, Cambridge University press, Cambridge, 1990.

[9]. R. M. Dickau, "Bell Number Diagrams."


http://mathforum.org/advanced/robertd/bell.html.
[9]. Dickau, R. M. "Stirling Numbers of the First Kind."
http://mathforum.org/advanced/robertd/stirling1.html.

[10]. L. E. Dikson, History of the theory of numbers, Vol. 1, Chelsea,


New York, NY, 1952.

[11]. T. Fort, Finite Differences and Difference Equations in the Real


Domain. Oxford, England: Clarendon Press, 1948.

[12]. I. Gessel and G. C. Rota, eds, Classic Papers in Combinatorics,


Birkhauser, Bostom, 1987.

[13]. R. L. Graham, D. E. Knuth, and O. Patashnik, Concrete


Mathematics: A Foundation for Computer Science, 2nd ed. Reading,
MA: Addison-Wesley, 1994.

[14]. G. H. Hardy, and Ramanujan: Twelve Lectures on Subjects


Suggested by His Life and Work, 3rd ed. New York: Chelsea, , 1999.

[15]. F. T. Horward, Congruences for the Stirling numbers and


Associated Stirling Numbers, Acta Atith. 55 (1990), 29 - 41.

[16]. Joe De Maio, Stirling Numbers of the Second kind and Primality,
Proceedings of the 2008 International Conference on Foundation of
Computer Science, pp. 124 – 130, (2008).

[17]. C. Jordan, Calculus of Finite differences, Chelsea, New York,


1960.

[18]. D. E. Knuth, The Art of Computer Programming, Vol I. Addison-


Wesley, Reading Mass (1968).

[19]. Koepf, W. Hypergeometric Summation: An Algorithmic Approach


to Summation and Special Function Identities. Braunschweig,
Germany: Vieweg, 1998.

[20]. J. l. Lagrange, Euvres iii. 425, Nouveaux memoires de l’


Academic royale de Berlin 2. (1773), p. 125.
[21]. J. Peterson, Beviser for Wilsons og Fermats Theorem, Tidsskrift
for mathematic 2 (1872), 64 – 65.

[22]. J. Riordan, An Introduction to Combinatorial Analysis, Wiley, New


York, 1968.

[23]. Roberto Sanchez – Peregrino, The Lucas Congruence for Stirling


numbers of the second kind, Acta Arithmetica, XCIV. 1 (2000). 41 – 51.

[23]. S. Roman, The Umbral Calculus. New York: Academic Press,


1984.

[24]. S. Roman, "The Lower Factorial Polynomial." §1.2 in The Umbral


Calculus. New York: Academic Press, pp. 5, 28-29, and 56-63, 1984

[25]. G. C. Rota, On the foundations of combinatorial theory, I:


Theory of Möbius functions, Z. Wahrsch. Theorie 2 (1964), 340 – 368.

[26]. G. C. Rota and B. E. Sagan, Congruences derived from group


action, European J. Combin. 1 (1980), 67 - 76.

[27]. H. J. Ryser, Combinatorial Mathematics, C arus Mathematical


Monograph No. 14. Math. Ass. Of America, Washington D. C. 1973.

[28]. B. E. Sagan, Congruences via Abelian Groups, Journal of


Number Theory, vol. 20 (1985), 210 – 237.

[29]. J. Touchard, Properties of arithmetic ques de certains numbers


recurrent, Ann. Soc. Sci. Bruxelles, Vol. 53 ( 1933), 21 – 31.

S-ar putea să vă placă și