Jump to content

Stirling number

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 80.201.75.22 (talk) at 19:45, 6 January 2006. The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

In mathematics, Stirling numbers arise in a variety of combinatorics problems. They are named after James Stirling, who introduced them in the 18th century. Two different sets of numbers bear this name: the Stirling numbers of the first kind and the Stirling numbers of the second kind.

Notation

Several different notations for the Stirling numbers are in use. Stirling numbers of the first kind are written with a small s, and those of the second kind with a large S:

The notation of using brackets and braces, in analogy to the binomial coefficients, was introduced in 1935 by Jovan Karamata and promoted later by Donald Knuth; it is referred to as Karamata notation.

Stirling numbers of the first kind

In combinatorial mathematics, unsigned Stirling numbers of the first kind

(with a lower-case "s") count the number of permutations of n elements with k disjoint cycles.

Stirling numbers of the first kind (without the qualifying adjective unsigned) are the coefficients in the expansion

where (x)n is the rising factorial

Stirling numbers of the first kind are sometimes written with the alternate notation

The definition can be inverted to express the falling factorial as a power series:

Similar relationships involving the Stirling numbers hold for the Bernoulli polynomials. Many relations for the Stirling numbers shadow similar relations on the binomial coefficients. The study of these 'shadow relationships' is termed umbral calculus and culminates in the theory of Sheffer sequences.

Table of values

Below is a table of values for the Stirling numbers of the first kind, similar in form to Pascal's triangle:

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
9 0 40320 -109584 118124 -67284 22449 -4536 546 -36 1

Recurrence relation

The Stirling numbers of the first kind obey the recurrence relation

for , with the initial conditions

The above follows from the recurrence relation on the falling factorials:

Simple identities

Note that

and

and

Other relations include

and where is a harmonic number and

where is a generalized harmonic number. A generalization of this relation th harmonic numbers is given in a later section.

Generating function

A variety of identities may be derived by manipulating the generating function

In particular, the order of summation may be exchanged, and derivatives taken, and then t or x may be fixed.

Finite sums

A simple sum is

Infinite sums

Some infinite sums include

which holds for .

Relation to harmonic numbers

Stirling numbers of the first kind can be expressed in terms of the harmonic numbers as follows:

where and

In the above, is the Gamma function and is the harmonic number.

Enumerative interpretation.

The absolute value of the Stirling number of the first kind, , counts the number of permutations of objects with exactly orbits (equivalently, with exactly cycles). For example, , corresponds to the fact that the symmetric group on 4 objects has permutations of the form

— 2 orbits of size 2 each

and permutations of the form

— 1 orbit of size 3, and 1 orbit of size 1

(see the entry on cycle notation for the meaning of the above expressions.)

Let us prove this. First, we can remark that the unsigned Stirling numbers of the first are characterized by the following recurrence relation:

To see why the above recurrence relation matches the count of permutations with cycles, consider forming a permutation of objects from a permutation of objects by adding a distinguished object. There are exactly two ways in which this can be accomplished. We could do this by forming a singleton cycle, i.e. leaving the extra object alone. This accounts for the term in the recurrence formula. We could also insert the new object into one of the existing cycles. Consider an arbitrary permutation of object with cycles, and label the objects , so that the permutation is represented by

To form a new permutation of objects and cycles one must insert the new object into this array. There are, evidently ways to perform this insertion. This explains the term of the recurrence relation. Q.E.D.

Stirling numbers of the second kind

Stirling numbers of the second kind S(n,k) (with a capital "S") count the number of equivalence relations having k equivalence classes defined on a set with n elements. The sum

is the nth Bell number. If we let

(in particular, (x)0 = 1 because it is an empty product) be the falling factorial, we can characterize the Stirling numbers of the second kind by

(Confusingly, the notation that combinatorialists use for falling factorials coincides with the notation used in special functions for rising factorials; see Pochhammer symbol.)

Table of values

Below is a table of values for the Stirling numbers of the second kind:

n \ k 0 1 2 3 4 5 6 7 8 9
0 1
1 0 1
2 0 1 1
3 0 1 3 1
4 0 1 7 6 1
5 0 1 15 25 10 1
6 0 1 31 90 65 15 1
7 0 1 63 301 350 140 21 1
8 0 1 127 966 1701 1050 266 28 1
9 0 1 255 3025 7770 6951 2646 462 36 1

Recurrence relation

These obey the recurrence relationship

with

.

Simple identities

Some simple identities include

This is because dividing n elements into n − 1 sets means each set wouldn't be larger than 2 (otherwise, there will be only n − 3 elements we have to put into at least n − 2 sets, each set contains at least one element) actually due to the same logic there can be exactly one set whose size is 2. Therefore we actually have to pick the unique two elements which would be in this set. Thus the above formula;

and

This is since dividing n elements into two sets A and B, allowing one set to be empty is 2n (each element has 2 possiblities to select a set to be in). Now we need to remove two unallowed states one where A is empty and the other where B is empty. Except since we counted each division of states where some elements are in A an all the others in B and the case where all the others were in A and those who were in A are in B also they are interepreted as a single case in our set division, we must divide 2n − 2 by 2, and thus we get the formula above.

Explicit formula

The Stirling numbers of the second kind are given by the explicit formula

.

Relation to Poisson distribution

The Stirling numbers of the second kind enjoy the following relationship with the Poisson distribution: if X is a random variable with a Poisson distribution with expected value λ, then its nth moment is

In particular, the nth moment of the Poisson distribution with expected value 1 is precisely the number of partitions of a set of size n, i.e., it is the nth Bell number (this fact is "Dobinski's formula").

Inversion relationships

The Stirling numbers of the first and second kind can be considered to be inverses of one-another:

and

where is the Kronecker delta.

See also

References

  • D.E. Knuth, Two notes on notation (TeX source).
  • Louis Comtet Advanced Combinatorics: The Art of Finite and Infinite Expansions, Reidel Publishing Company, Dordrecht-Holland/Boston-U.S.A., 1974.
  • "Stirling numbers of the first kind". PlanetMath.
  • "Stirling numbers of the second kind". PlanetMath.
  • Victor Adamchick, On Stirling Numbers and Euler Sums, (1996)
  • Arthur T. Benjamin, Gregory O. Preston, Jennifer J. Quinn, A Stirling Encounter with Harmonic Numbers, (2002) Mathematics Magazine, 75 (2) pp 95-103.
  • J. M. Sixdeniers, K. A. Penson, A. I. Solomon, Extended Bell and Stirling Numbers From Hypergeometric Exponentiation (2001), Journal of Integer Sequences, 4, Article 01.1.4.
  • Hsien-Kuei Hwang, Asymptotic Expansions for the Stirling Numbers of the First Kind (1994).

Stirling numbers of the first kind at PlanetMath.